7 Floristic Relationships of Seasonally Dry Forests of Eastern South

Transcrição

7 Floristic Relationships of Seasonally Dry Forests of Eastern South
2987_C007.fm Page 151 Thursday, December 1, 2005 7:03 PM
Relationships of
7 Floristic
Seasonally Dry Forests of Eastern
South America Based on Tree
Species Distribution Patterns
Ary T. Oliveira-Filho, João André Jarenkow, Maria Jesus
Nogueira Rodal
CONTENTS
7.1
7.2
Introduction...........................................................................................................................152
Methods ................................................................................................................................154
7.2.1 Preparation and Revision of the Databases .............................................................154
7.2.2 Vegetation Classification ..........................................................................................157
7.2.3 Multivariate Analyses ...............................................................................................158
7.2.4 Condensed Floristic Data .........................................................................................159
7.3 Results...................................................................................................................................160
7.3.1 Multivariate Analyses ...............................................................................................160
7.3.2 Analyses of Condensed Floristic Information .........................................................162
7.4 Discussion.............................................................................................................................170
7.5 Conclusion ............................................................................................................................175
Acknowledgements ........................................................................................................................176
References ......................................................................................................................................176
Appendix. Most Frequent Species (>70% of Checklists) in the Tree
flora of Selected SDTF and SDSF Formations of Eastern South America..................................179
ABSTRACT
The tree flora of seasonally dry forests (SDTF) of eastern tropical and subtropical South America
was investigated according to two main aspects: (a) the variations in floristic composition were
analyzed in terms of geographical and climatic variables by performing multivariate analyses on
532 existing floristic checklists; and (b) the links among different seasonally dry forest formations,
Amazonian forests and cerrados (woody savannas) were assessed. Analyses were performed at the
species, genus and family levels. There was a strong spatial pattern in tree species distribution that
only receded and allowed clearer climate-related patterns to arise when either the geographical
range was restricted or data were treated at the genus and family levels. Consistent floristic
differences occurred between rain and seasonal forests, although these were obscured by strong
regional similarities which made the two foresttypes from the same region closer to each other
floristically than they were to their equivalents in different regions. Atlantic rain and seasonal forests
151
2987_C007.fm Page 152 Thursday, December 1, 2005 7:03 PM
152
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
were floristically closer to each other than to Amazonian rain forests but north-east rain and seasonal
forests were both closer to Amazonian rain forests than each other, though only at the generic and
familial levels. Atlantic seasonal forests also share a variable proportion of species with caatingas,
cerrados and the chaco, and may represent a transition to these open formations. Increasing periods
of water shortage, with increases in soil fertility and temperature are characteristic of a transition
from semideciduous to deciduous forests and then to the semi-arid formations, either caatingas
(tropical) or chaco forests (subtropical), while increasing fire frequency and decreasing soil fertility
lead from seasonal forests to either cerrados (tropical) or southern campos (subtropical). The SDTF
vegetation of eastern South America may be classified into three floristic nuclei: caatinga, chaco
and Atlantic forest (sensu latissimo). Only the last, however, should be linked consistently to the
residual Pleistocenic dry seasonal flora (RPDS). Caatinga and chaco represent the extremes of
floristic dissimilarity among the three nuclei, also corresponding to the warm-dry and warm-cool
climatic extremes, respectively. In contrast to the caatinga and chaco nuclei, the Atlantic SDTF
nucleus is poor in endemic species and is actually a floristic bridge connecting the two drier nuclei
to rain forests. Additionally, there are few grounds to recognize the Atlantic nucleus flora as a
clearly distinct species assemblage, since there is a striking variation in species composition found
throughout its wide geographical range. Nevertheless, there is a group of wide-range species that
are found in most regions of the Atlantic nucleus, some of which are also part of the species blend
of the Caatinga and Chaco floras, though the latter plays a much smaller part. We propose that it
is precisely this small fraction of the Atlantic nucleus flora that should be identified with the RPDS
vegetation.
7.1 INTRODUCTION
Neotropical seasonally dry tropical forests, or SDTF, are presently an increasing focus of attention
because of both their very threatened status and poorly studied flora and ecology, and this is striking
when compared to traditional flag ecosystems like rain forests and savannas (Mooney et al., 1995).
They occur where annual rainfall is less than 1600 mm and more than 5–6 months receive less than
100 mm (Graham and Dilcher, 1995) and therefore include a diverse array of vegetation formations,
from tall semideciduous forests to thorny woodland with succulents (Murphy and Lugo, 1995).
Despite all this variation, Pennington et al. (2000) argue that the concept of SDTF should exclude
fire-related formations, such as savannas and cerrados, and the non-tropical chaco forests.
The distribution of SDTF in South America forms an arc with the ends positioned at the caatinga
domain of north-eastern Brazil and the Caribbean coast of Colombia and Venezuela and a long
curved route connecting the ends through the seasonal forests of the Atlantic forest domain, the
patches of seasonal forests of the cerrado domain, and the seasonal forests of the Andean piedmont,
inter-Andean valleys, Pacific coast and Caribbean coast. Prado (1991) and Prado and Gibbs (1993)
suggested the hypothesis that this arc is a relic of a much wider distribution of SDTF in South
America reached during the Pleistocene glacial maxima. They based their model on the present
distribution of what they called residual Pleistocenic dry seasonal (RPDS) flora. Since then a number
of studies have analysed species distribution patterns of this flora in order to assess the validity of
the RPDS arc hypothesis in different geographical contexts (e.g. Pennington et al., 2000, 2004;
Prado, 2000; Bridgewater et al., 2003; Linares-Palomino et al., 2003; Spichiger et al., 2004). The
assessment of floristic links among species assemblages of SDTF areas scattered over the putative
RPDS arc has proved a useful tool to elucidate patterns of historical vegetation change. LinaresPalomino et al. (2003) performed a detailed phytogeographical analysis of SDTF areas of Pacific
South America, i.e. the western section of the RPDS Arc, and found three main groups with a
considerable dissimilarity among them. In the present contribution, we perform a similar analysis
of seasonally dry forest areas of the eastern section of the RPDS arc. As we dealt with both the
tropical and subtropical regions of eastern South America we incorporated seasonally dry subtropical forests into the SDTF concept.
2987_C007.fm Page 153 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
Atlantic forest formations
Tropical rain forest
Tropical seasonal semideciduous forest
Tropical seasonal deciduous forest
Subtropical rain forest
Subtropical araucaria rain forest
Subtropical seasonal semideciduous forest
Subtropical seasonal deciduous forest
153
4°
Caatingas
Ca
Cer
rado
s
at
in
ga
8°
s
12°
16°
Cerrados
Cerrados
20°
Chaco
24°
Chaco
n
ea
ti
lan
At
c
cO
28°
pos
Cam
32°
68°
64°
60°
56°
52°
48°
44°
40°
36°
32°
FIGURE 7.1 Map of eastern South America showing the distribution of the predominant vegetation formations
of the South American Atlantic forest domain. Caatingas, cerrados, chaco and campos are the adjacent domains
that make up the “diagonal of open formations”.
The geographical range of the SDTF areas analysed in the present study is large enough to
include four vast vegetation domains: Atlantic Forest, caatinga, cerrado and chaco (Figure 7.1).
The Atlantic forest domain stretches for >3300 km along the eastern Brazilian coast between the
latitudes of 6°S and 30°S and makes up the second largest tropical moist forest area of South
America, exceeded only by the vast Amazonian domain. The two forest domains are separated by
the so-called diagonal of open formations, a corridor of seasonally dry formations that includes
another three domains: the caatingas (mostly tropical thorny woodlands), cerrado (mostly woody
savannas), and the chaco (mostly subtropical thorny woodlands). Each domain contains its particular
SDTF formations. The now widely accepted concept of Atlantic forests (sensu lato) attaches
seasonal forests, the Atlantic SDTF, to the coastal rain forests, formerly considered as the true
(sensu stricto) Atlantic forests (Oliveira-Filho and Fontes, 2000; Galindo-Leal and Câmara, 2003).
Caatingas and carrascos (tropical deciduous scrubs) are both SDTF and make up the predominant
vegetation cover of the caatinga domain. SDTF formations are also an important component of the
cerrado domain where they occur as forest patches on more fertile soils and on the freely drained
slopes of gallery forests (Oliveira-Filho and Ratter, 1995, 2002). In the chaco domain, SDTF occur
on peripheral areas and in some internal forest patches (Prado, 2000).
2987_C007.fm Page 154 Thursday, December 1, 2005 7:03 PM
154
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Atlantic SDTF occur as seasonal (semideciduous and deciduous) forests all along the contact
zone between rain forests and the diagonal of open formations, comprising three different scenarios
(see Figure 7.1 for distribution and Table 7.1 for nomenclature). (a) In north-eastern Brazil, SDTF
form a narrow belt (<50 km) in the sharp transition between coastal rain forests and the semiarid
caatingas, but also occur as hinterland montane forest enclaves, the brejos (Rodal, 2002; Rodal and
Nascimento, 2002). (b) The transition between coastal rain forests and cerrados in south-eastern
Brazil involves a much larger extent of SDTF that becomes increasingly wider towards the south
to reach eastern Paraguay and north-eastern Argentina. They also form complex mosaics with
cerrado vegetation to the west so that if the SDTF component of the cerrado domain is seen as an
extension of the Atlantic SDTF, as proposed by Oliveira-Filho and Ratter (1995), a concept of
Atlantic forests sensu latissimo must be created. (c) In the southern subtropical realm, large extents
of hinterland araucaria rain forests are attached to the coastal subtropical rain forests, and SDTF
appear in the west and south as transitions to both chaco forests and southern campos, or pampa
prairies (Spichiger et al., 1995; Quadros and Pillar, 2002).
In the present study we sought patterns of floristic differentiation among SDTF areas of eastern
tropical and subtropical South America that could be associated with geographical and climatic
variables, and assessed the floristic links of seasonally dry forest formations of different regions,
Amazonian forests and cerrados. We addressed the following questions: (a) How strongly differentiated are Atlantic seasonal and rain forests in different sections of their geographical range?
(b) To what extent is the tree species composition of Atlantic seasonal forests transitional between
those of rain forests and open formations, such as caatingas and cerrados? (c) Is the Atlantic rain
forest flora closer to that of Amazonian rain forests or to that of the Atlantic seasonal forest?
(d) How strong are the floristic links among caatingas, the seasonal forests of the Atlantic and cerrado
domains, and the chaco forests? (e) Does SDTF flora change its composition in response to climatic
variations? and (f) How are the above questions answered at the species, genus and family levels?
7.2 METHODS
7.2.1 PREPARATION
AND
REVISION
OF THE
DATABASES
We selected from the literature a total of 659 papers containing floristic checklists produced by
surveys of the tree component of 532 areas of eastern tropical and subtropical South America. The
geographical range included the Atlantic forest, caatinga, cerrado and chaco domains (Figure 7.1).
Vegetation formations included seasonally dry tropical forests (SDTF), which are the focus of the
present study, as well as tropical and subtropical rain forests, and subtropical araucaria rain forests
(see Table 7.1). SDTF formations (Figures 7.2 and 7.3) are a broad category that contains tropical
and subtropical seasonal forests (both deciduous and semideciduous) as well as caatingas and
carrascos. Mesotrophic cerradões were treated as SDTF-cerrado transition.
Individual areas were defined arbitrarily within a maximum range of 20 km width and 400 m
elevation, and thus included sections of large continuous forest tracts (e.g. Tiradentes), assemblages
of forest fragments (e.g. Santa Maria) and nearby areas at different altitudes (e.g. Lençóis and
Palmeiras). We obtained the following geographical information for each area: latitude and longitude at the centre of the area, median altitude, and shortest distance from the ocean. We also
obtained the annual and monthly means for the temperature and rainfall of each area or the nearest
meteorological stations. When the source of the checklist did not provide the climatic records, they were
obtained from DNMet (1992) and from governmental websites (http://masrv54.agricultura.gov.br/rna;
http://www.inmet.gov.br/climatologia). Some areas required interpolation and/or standard correction for altitude (Thornthwaite, 1948).
We entered the information from the 532 areas on to spreadsheets using Microsoft Excel 2002
in order to produce two databases. The first consisted of basic information about each area including
locality, forest classification (see below), geographical and climatic variables, and literature sources.
lowland
submontane
lower montane
upper montane
lowland
submontane
lower montane
lowland
submontane
lower montane
lowland
submontane
lower montane
upper montane
lowland and
submontane
lowland to
lower montane
lowland to lower
montane
lowland
lowland
submontane
lower montane
upper montane
lowland
submontane
lower montane
upper montane
lower montane
upper montane
low altitude
low-high
altitude
low altitude
low altitude
high altitude
low altitude
high altitude
low altitude
high altitude
low altitude
high altitude
low altitude
high altitude
high altitude
low altitude
high altitude
low altitude
Altitudinal Belt
South
South-west
North-east and
east
Central-west
South and
south-west
South
North-east, east,
south-east and
central-west
North-east, east,
south-east and
central-west
South and
south-east
South
North-east,east
and south-east
Regions
Atlantic seasonal forests (S) and
Peripheral Chaco seasonal
forests (SW)
Atlantic seasonal forests (S)
Atlantic seasonal forests
(NE/E/SW) and Centralwestern seasonal forests (CW)
Atlantic rain forests
(Atlantic forests sensu stricto)
Main Formations
Atlantic forests
(Atlantic forests
sensu latissimo;
sensu lato excludes
CW and SW)
SDTF –
Seasonally dry
tropical forests
Rain forests
Floristic Relationships of Seasonally Dry Forests of Eastern
Southern campos or pampa prairies
Caatingas (tropical thorny woodlands)
Carrascos (tropical deciduous scrubs)
Cerrados (sensu lato: open savannas to
forests, or cerradões)
Chaco (subtropical thorny woodland)
Subtropical seasonal deciduous forests
Subtropical seasonal semideciduous forests
Tropical seasonal deciduous forests
Tropical seasonal semideciduous forests
Subtropical araucaria rain forests
Subtropical rain forests
Tropical rain forests
Vegetation Formations
TABLE 7.1
Nomenclature Used in the Present Chapter for Vegetation Classification of Eastern Tropical and Subtropical South America.
2987_C007.fm Page 155 Thursday, December 1, 2005 7:03 PM
155
2987_C007.fm Page 156 Thursday, December 1, 2005 7:03 PM
156
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
A
B
C
D
E
F
FIGURE 7.2 Seasonally dry tropical forests (SDTF) of eastern South America: (A) caatinga in São Raimundo
Nonato, Piauí; (B) submontane tropical seasonal deciduous forest in the Serra das Confusões, Piauí; (C–F)
submontane tropical seasonal deciduous forest in Três Marias, Minas Gerais in the dry (C and D) and wet
(E and F) seasons (Image credits: F. Filetto [A and B] and M. A. Fontes [C–F]).
The second database was a matrix of tree species presence in the 532 areas plus three additional
checklists that we included to compare them to Amazonian rain forests, cerrados (s.l.) and chaco
forests. The first of these combined the flora of Reserva Ducke (Ribeiro et al., 1999) with the
22 checklists of Amazonian rain forests compiled by Oliveira-Filho and Ratter (1994) and contained
2190 tree species. The second contained 528 species present in 98 areas of cerrado (Ratter et al.,
1996) and the third 183 chaco species listed by Prado (1991), Lewis et al. (1994) and Spichiger
et al. (1995).
Before reaching its final form, the information contained in the species database underwent a
detailed revision to check all species names cited in the checklists for growth form, synonymy and
geographical distribution. Only species capable of growing to trees or treelets, i.e. producing a freestanding woody stem >3 m in stature, were maintained in the database. The task involved consultation
of 387 published revisions of families and genera, 32 specialists of various institutions and four
websites (http://www.cnip.org.br; http://www.ipni.org/index.html, 2003-2005; http://www.mobot.org/
W3T/Search/vast.html; http://sciweb.nybg.org/science2/hcol/sebc/index.asp). When these sources
2987_C007.fm Page 157 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
A
C
157
B
D
FIGURE 7.3 Seasonally dry tropical forests (SDTF) of eastern tropical and subtropical South America:
(A) lower montane tropical semideciduous forest in Itambé do Mato Dentro, Minas Gerais; (B) submontane
tropical seasonal semideciduous forest in the Chapada dos Guimarães, Mato Grosso; (C) lowland subtropical
seasonal semideciduous forest in Praia do Tigre, Rio Grande do Sul; (D) lowland subtropical seasonal
deciduous forest in Cachoeira do Sul, Rio Grande do Sul (Image credits: A.T. Oliveira-Filho [A, B], J. A.
Jarenkow [D] and J. C. Budke [E]).
referred to herbarium specimens unequivocally collected in any of the 532 areas, the species was
added to the database. The final database contained 6598 species, 976 genera, and 128 families. The
species classification into families followed the Angiosperm Phylogeny Group II (APG, 2003).
7.2.2 VEGETATION CLASSIFICATION
We extend here the vegetation classification proposed by Oliveira-Filho and Fontes (2000) for
south-east Brazil to include a much wider geographical range as well as additional vegetation
formations (Table 7.1). This extended classification was based on exploratory multivariate analyses
of both floristic and climatic data (ongoing studies). We defined the top classification level by
combining main thermoclimate (either tropical or subtropical) and rainfall seasonality (rainy,
seasonally rainy and semi-arid) and established the limit between the two thermoclimates at the
latitudes of 25°30’S and 26°30’S for rain and seasonal forests, respectively. Areas of the chaco
domain and araucaria rain forests were all subtropical; areas of the cerrado and caatinga domains
were all tropical. We classified areas with tropical climates as rain forests, seasonal forests/cerrados
and caatingas/carrasco in which the dry season lasts for up to 30 days (rainy), > 30 − 160 days
(seasonally rainy) and >160 days (semi-arid), respectively. In subtropical climates, we classified
the areas as chaco forests/peripheral chaco seasonal forests where the dry season lasts for > 30 days
(semi-arid) and areas below this limit as either rain forests or seasonal forests/campos in which the
difference in mean monthly temperatures between the coolest and warmest months is up to 10°C
or > 10 − 15°C, respectively (rainfall seasonality secondary).
2987_C007.fm Page 158 Thursday, December 1, 2005 7:03 PM
158
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Seasonal deciduous and semideciduous forests of both tropical and subtropical climates are
commonly distinguished by the amount of leaffall during the slow-growth season (Veloso et al.,
1991). Except for the degree of deciduousness, it is often difficult to tell them apart, particularly
in central Brazil where they commonly form continua determined by local variations of soil moisture
and fertility (Oliveira-Filho and Ratter, 2002). Therefore, in most cases we opted to trust the authors’
experience to classify seasonal forests as either deciduous or semideciduous. Savannas, i.e. cerrado
(sensu lato) and campo are also a very important component of the vegetation in seasonal climates
but we included only the mesotrophic cerradão (plural, cerradões) in the analyses because of its
forest-like physiognomy. Again, the authors’ experience was trusted to distinguish this vegetation
from seasonal forests. Another distinction was made between subtropical araucaria rain forests and
other rain forests, based on their geographical location in the inner highlands and the conspicuous
presence of emergent trees of Araucaria angustifolia (Bert.) O.Kuntze. Areas of caatinga and
carrasco were distinguished by elevation and substrate, the former occurring on dissected lowlands
and the latter on sandy plateaus (see Queiroz, Chapter 6).
We also used the exploratory multivariate analyses of floristic data to classify the above
vegetation formations according to altitude and geographical region. We defined elevation ranges
as follows. For latitudes <16ºS: lowland, <400 m; submontane, 400 − <800 m; lower montane,
800 − <1200 m; upper montane, >1200 m. For latitudes between 16° and <23°30’S: lowland,
<300 m; submontane, 300 − <700 m; lower montane, 700 − <1100 m; upper montane, >1100 m.
For latitudes between 23º30’ and <32ºS, lowland, <200 m; submontane, 200 – <600 m; lower
montane, 600 − <1000 m; upper montane, >1000 m. The geographical regions recognized were
north-east, east, south-east, south, central-west and south-west. The resulting classification
categories are given in Table 7.1 and the geographical distribution of the 532 areas are shown
in Figure 7.4. Limited space does not allow us to list the areas, neither to provide here their
description and source references. We intend to make this information available in a forthcoming
publication.
7.2.3 MULTIVARIATE ANALYSES
We used detrended correspondence analysis, DCA (Hill and Gauch, 1980), processed by the
program PC-ORD 4.0 (McCune and Mefford, 1999) to seek main species distribution gradients
across 341 SDTF areas. We removed rain forests because the patterns within this vegetation
type were not the focus of the present work. An additional DCA was performed with the 243
areas of tropical seasonal forests (subtropical and semiarid formations excluded) to seek more
detailed patterns within the group. Two other DCA were performed separately for the areas of
caatinga and subtropical seasonal forests. We chose DCA coupled to a posteriori interpretation
of ordination results because we aimed at patterns dictated solely by the species without the
interference of environmental variables, as occurs with joint analyses such as CCA (Kent and
Coker, 1992). We used two interpretation tools: the previous vegetation classification of the
areas and 13 geographical and climatic (hereafter geo-climatic) variables. They were both
plotted (a posteriori) on the DCA diagrams as symbols and arrows, respectively. The geoclimatic variables were latitude, longitude, median altitude, distance from the ocean, mean
annual temperature, mean monthly temperatures in the warmest and coolest months, mean
temperature range obtained from the difference between the two previous variables, mean
annual rainfall, mean monthly rainfall of the dry (June–August) and rainy (December–February)
seasons, rainfall distribution ratio obtained from the proportion between the two previous
variables and mean duration of the dry season obtained from the number of days of water
shortage given by Walter diagrams (Walter, 1985). We also obtained the Pearson correlation
coefficients between the geo-climatic variables and the ordination scores of the areas in each
DCA axis.
2987_C007.fm Page 159 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
159
2
Vegetation classification
Lowland rain forest
Submontane rain forest
Lower montane rain forest
Upper montane rain forest
Lower montane araucaria rain forest
Upper montane araucaria rain forest
Lowland seasonal semideciduous forest
Submontane seasonal semideciduous forest
Lower montane seasonal semideciduous forest
Upper montane seasonal semideciduous forest
Lowland seasonal deciduous forest
Submontane seasonal deciduous forest
Lower montane seasonal deciduous forest
Upper montane seasonal deciduous forest
Caatinga
Carrasco
Mesotrophic cerradão
Northeast
4
6
8
10
12
14
16
Central-West
18
20
Latitude (S)X
East
22
24
Southeast
Southwest
26
28
30
South
32
34
66
64
62
60
58
56
54
52 50 48 46
Longitude (W)X
44
42
40
38
36
34
32
FIGURE 7.4 Geographical coordinates diagram showing the location and vegetation classification of the 532
areas used in the analyses and the six geographical regions. Forest areas situated in the north-east, east, southeast and central-west regions are classified as tropical and those in the south and south-west are subtropical.
7.2.4 CONDENSED FLORISTIC DATA
Because both previous and present multivariate analyses demonstrated that the vegetation classification system adopted was highly consistent, we eventually condensed the floristic information
contained in the database by lumping together the species records within main vegetation formations
(Table 7.1). Atlantic rain forests as well as the cerrado, chaco and Amazonian rain forest checklists
were incorporated here. We also merged lowland and submontane categories as low altitude and
lower and upper montane categories as high altitude. The resulting lumped matrix of binary data
of species presence in the main vegetation formations was used to produce two additional matrices,
for genera and families. The generic and familial matrices were both quantitative, as they consisted
of the number of species per genus or family, respectively, in each main vegetation formation. We
performed cluster analyses of the condensed matrices using the program PC-ORD 4.0. Cluster
analyses used Jaccard’s floristic similarity for species and relative squared Euclidian distances for
genera and families (number of species as abundance data); the linkage method was group average
(Kent and Coker, 1992). We also used the condensed data to perform a direct quantitative assessment
of the floristic links between the vegetation formations by plotting the number of shared and
exclusive species in Venn diagrams. The most frequent species, and the richest genera and families
of main vegetation formations were extracted from the matrices.
2987_C007.fm Page 160 Thursday, December 1, 2005 7:03 PM
160
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
7.3 RESULTS
7.3.1 MULTIVARIATE ANALYSES
The ordination diagrams yielded by DCA are shown in Figures 7.5 and 7.6 for the two assemblages
of vegetation areas, seasonally dry tropical forests (SDTF) and tropical seasonal forests. Their
eigenvalues are first DCA, 0.688 (axis 1) and 0.394 (axis 2); second DCA, 0.400 (axis 1) and 0.475
(axis 2). According to ter Braak (1995), these eigenvalues are relatively high (>0.3), indicating
considerable species turnover along the gradients summarized in the first two axes. In addition,
most DCA axes produce a number of high values of Pearson correlation coefficients between geoclimatic variables and ordination scores (Table 7.2) giving consistency to the interpretation of the
emerging patterns.
The first ordination axis in the DCA for SDTF is chiefly correlated with latitude, minimum
monthly temperature, duration of the dry season, annual temperature and annual rainfall (Table 7.2).
This indicates that the data structure summarized by the first axis primarily reflects a geographical
gradient based on latitude which corresponds to a major climatic gradient towards the south characterized by decreasing temperatures and duration of the dry season and increasing total rainfall.
Longitude and distance to the ocean are more strongly correlated with the second DCA axis but no
climatic variable accompanies this gradient. The areas of caatinga and carrasco are found at the rightside of the diagram associated with latitudes near the Equator, longer dry seasons and higher temperatures (Figure 7.5). No distinction is made between north-east and east caatingas but the three carrasco
areas are displaced to the top on the second axis together with three areas of caatinga (Serra da
Capivara, São José do Piauí and Ibiraba) which differ from other caatingas in their sandy substrate,
80
Longitude
D Ocean
Axis 2
60
T annual
Latitude
40
T mth min
Dry season
T range
R annual
20
0
Caatingas:
NE
E
Carrascos:
NE
Mesotrophic cerradão
CW
0
Tropical semideciduous forests:
NE
SE
CW
E
Tropical deciduous forests:
NE
CW
Subtropical seasonal Forests:
S
SW
40
80
Axis 1
FIGURE 7.5 Diagram yielded by detrended correspondence analysis (DCA) showing the ordination of 341
areas of seasonally dry tropical forests (SDTF) of eastern South America on the first two DCA axes, based
on the presence of 3018 tree species. The areas are classified according to main vegetation formation and
geographical region. The centred straight-lines show the correlation between axes and geoclimatic variables
(only those with r > 0.3 with at least one axis are shown): T = temperature, Mth = monthly, Min = minimum,
R = rainfall, D = distance.
2987_C007.fm Page 161 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
161
TABLE 7.2
Detrended Correspondence Analysis (DCA)
Vegetation Physiognomies and DCA Axes
Geoclimatic variables
Latitude
Longitude
Altitude
Distance to the ocean
Annual temperature
Minimum monthly temperature
Maximum monthly temperature
Monthly temperature range
Annual rainfall
Monthly rainfall in JJA
Monthly rainfall in DJF
Rainfall distribution ratio
Duration of the dry season
SDTF and SDSF
Tropical Seasonal Forests
(N = 341 areas)
(N = 243 areas)
Axis 1
Axis 2
Axis 1
Axis 2
–0.88
–0.52
–0.14
0.12
0.77
0.81
0.48
–0.66
–0.72
–0.27
–0.52
0.08
0.79
0.19
0.72
–0.26
0.66
0.15
0.08
0.26
0.14
0.10
–0.17
0.21
–0.30
0.06
0.08
–0.65
0.58
–0.73
–0.63
–0.54
–0.59
0.15
0.08
0.05
0.07
0.08
–0.07
–0.91
–0.55
–0.13
–0.07
0.49
0.61
0.23
–0.62
–0.37
0.39
–0.67
0.60
0.41
Pearson correlation coefficients between geo-climatic variables and the ordination
scores of N areas of seasonally dry forests of the South American Atlantic forest
domain. Coefficients are given for the first two axes of DCAs performed for species
presence in three different sets of areas. Correlations >0.5 are in bold.
as do the carrascos. Towards the left-side of the diagram, caatingas and carrascos are followed by
areas of tropical seasonal forests of the north-east and central-west, discriminated at the bottom
and top halves of the diagram, respectively. For the north-east areas, deciduous forests come first
followed by semideciduous forests. For the central-west, however, deciduous and semideciduous
forests are not distinguished from each other and only mesotrophic cerradões form a consistent
clump. Along the latitudinal sequence of the first DCA axis, north-east tropical seasonal forests
are followed by those of the east and then south-east regions. This sequence ends at the left-side
of the diagram where the areas of subtropical seasonal forests lie that correspond to the extremes
of higher latitudes, lower temperatures and shorter dry seasons are situated. In addition, the second
axis discriminates, at the top, the four south-west areas of peripheral chaco seasonal forests.
The DCA performed for tropical seasonal forests shows additional patterns linked to altitude that
do not arise when other seasonally dry forests are included. The first axis is primarily correlated with
distance to the ocean, longitude, temperatures (annual, minimum and maximum) and altitude, while
the second axis is more strongly correlated with latitude, monthly rainfall in DJF, rainfall distribution
ratio and monthly temperature range (Table 7.2). This suggests that the first axis primarily reflects a
geographical gradient based on penetration into the continental interior together with decreasing
altitude, both of which also correspond to a climatic gradient characterized mainly by increasing
temperatures. To a great extent, the second axis repeats patterns already shown by the first axis of
the previous DCA, so that all north-east seasonal forests are strongly discriminated at the top of the
diagram (Figure 7.6). As opposed to the others, north-east areas show stronger correlations with
decreasing latitude, rainfall in DJF and temperature range and with increasing rainfall distribution
ratio. The first ordination axis, however, discriminates north-east areas of deciduous and semideciduous forests to the left- and right-sides of the diagram, respectively, with the single exception of the
oceanic island of Fernando de Noronha, located at the top right corner. At the bottom half of the
2987_C007.fm Page 162 Thursday, December 1, 2005 7:03 PM
162
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Axis 2
70
Tropical forest formations:
Semideciduous-NE
Semideciduous-E
Semideciduous-SE
Semideciduous-CW
Deciduous-CW
Deciduous-NE
T mth min
R D-ratio
T annual
50
T mth max
Altitude
D Ocean
Longitude
30
T range
Altitudinal range:
Lowland
Submontane
Lower montane
Upper montane
20
RD JF
40
60
Axis 1
80
Latitude
FIGURE 7.6 Diagram yielded by detrended correspondence analysis (DCA) showing the ordination of 243
areas of tropical seasonal forests of the South America Atlantic forest domain on the first two DCA axes,
based on the presence of 2680 tree species. The areas are classified according to forest formation, geographical
region and altitudinal range. The centred straight-lines show the correlation between axes and geoclimatic
variables (only those with r > 0.3 with at least one axis are shown): T = temperature, Mth = monthly, Min =
minimum, Max = maximum, D = distance, R = rainfall, DJF = December-January-February, D-Ratio =
distribution ratio.
diagram the first ordination axis discriminates central-west seasonal forests from those of the southeast and east regions in such a way that two concurrent geographical gradients were distinguished,
one related to longitude and distance from the ocean and the other with altitude, both involving
decreasing temperatures. The areas at the extreme left of the diagram are the westernmost low-altitude
seasonal forests of Mato Grosso and Bolivia, while those at the extreme right are mostly the eastern
high-altitude seasonal forests of Bahia and Minas Gerais. As in the previous DCA, central-west
deciduous and semideciduous forests are not discriminated from each other.
Additional DCAs were performed separately for the areas of caatinga and subtropical seasonal
forests, but no relevant additional patterns arose. Deciduous and semideciduous subtropical forests
of the south were not discriminated amongst themselves and only altitude showed a weak correlation
with the floristic patterns.
7.3.2 ANALYSES
OF
CONDENSED FLORISTIC INFORMATION
As they were extracted from floristic checklists for specific forest areas, the condensed information
must be regarded as a means of assessing the floristic links between the main forest formations
quantitatively and not as a register of actual figures for number of species, either total or in common.
The classification dendrograms (Figure 7.7) show different patterns for each of the three
taxonomic levels. A clear general trend arising from the species dendrogram is that regional patterns
2987_C007.fm Page 163 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
163
Information remaining (%)
100
TR-L-NE
TS-L-NE
TS-H-NE
TR-H-NE
TD-L-NE
Car-NE
Caa-NE
Caa-E
TR-L-E
TR-L-SE
TR-H-SE
TS-L-E
TS-H-E
TS-L-SE
TS-H-SE
TS-L-CW
TS-H-CW
TD-H-CW
TD-L-CW
SR-L-S
SR-H-S
SA-H-S
SA-H-SE
SS-L-S
SD-L-S
Cdm-CW
Cerr-CW
SS-H-S
SD-H-S
Amz
SD-L-SW
SD-H-SW
Cha-SW
TR-L-NE
TS-L-NE
Amz
TR-H-NE
TR-L-E
TS-L-E
TS-L-SE
TS-H-SE
TS-H-E
TS-L-CW
TS-H-CW
TR-L-SE
TR-H-SE
SR-L-S
SR-H-S
SA-H-S
SA-H-SE
SS-L-S
SD-L-S
SS-H-S
SD-H-S
TS-H-NE
TD-L-NE
TD-L-CW
TD-H-CW
Cdm-CW
Cerr-CW
Car-NE
Caa-NE
Caa-E
SD-L-SW
SD-H-SW
Cha-SW
TR-L-NE
TS-L-NE
Amz
TS-H-NE
TD-L-NE
TD-L-CW
TD-H-CW
Cdm-CW
SD-L-SW
SD-H-SW
Car-NE
Caa-NE
Caa-E
Cha-SW
TR-L-E
TS-L-E
TS-L-SE
TS-L-CW
TS-H-CW
TS-H-E
TS-H-SE
Cerr-CW
TR-H-NE
TR-L-SE
TR-H-SE
SR-H-S
SR-L-S
SA-H-S
SA-H-SE
SS-L-S
SD-L-S
SD-H-S
SS-H-S
75
50
25
0
Species
Genera
Families
FIGURE 7.7 Dendrograms produced by group averaging of Jaccard’s floristic similarity for species and relative squared
Euclidian distances for genera and families of the tree flora 23 areas of eastern Amazonian rain forests (Amz), 39 areas of
caatinga (Caa), 3 areas carrasco (Car), 376 areas of cerrado (Cerr), 11 areas of mesotrophic cerradão (Cdm), 5 areas of
chaco forests (Cha) and 479 areas of Atlantic forests s.l. merged into 28 main forest formations abbreviated as follows: the
first set of letters stands for either tropical (T) or subtropical (S) forests and for rain (R), araucaria rain (A), seasonal
semideciduous (S) and seasonal deciduous (D) forests; the middle letters stand for low (L) and high (H) altitude; the last
letters stand for north-east (NE), east (E), south-east (SE), central-west (CW), south (S) and south-west (SW) regions.
Scale in dendrograms expresses the remaining information after clustering.
2987_C007.fm Page 164 Thursday, December 1, 2005 7:03 PM
164
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
are stronger than vegetation formation patterns. Four main geographical groups are discriminated.
The first contains all vegetation formations of the north-east region, including tropical rain and
seasonal forests, caatingas and carrascos. The second, and largest, group contains tropical rain and
seasonal forests of four regions (east, south-east, central-west and south) plus cerrados and cerradões. The third main group is composed of the Amazonian rain forests and the fourth, and most
distinct, by chaco forests and peripheral chaco seasonal forests. The north-east main group is split
into two subgroups, the first containing moister formations (tropical rain forests and seasonal
semideciduous forests) and the second drier formations (tropical seasonal deciduous forests, caatingas and carrascos). The second main group is split into five subgroups, all of clear geographical
nature: (a) tropical rain and seasonal semideciduous forests of the east and south-east; (b) tropical
seasonal semideciduous and deciduous forests of the central-west; (c) subtropical low-altitude
seasonal forests and rain forests of the south, plus subtropical high-altitude araucaria rain
forests of both the south and south-east; (d) cerrados (sensu lato) and mesotrophic cerradões; and
(e) subtropical high-altitude seasonal forests of the south.
The dendrogram for genera shows different main groups and an increased role of vegetation
formation over regional patterns. The first main group contains tropical low-altitude rain forests
and semideciduous forests of the north-east and Amazonian regions. The second main group
contains two subgroups: (a) tropical seasonal semideciduous forests of the east, south-east and
central-west, plus tropical rain forests of the east, and (b) tropical rain forests of the south-east,
subtropical rain forests of the South and subtropical Araucaria rain forests of both the south and
south-east. The third, fourth and fifth main groups contain, respectively, subtropical seasonal forests
of the south, tropical seasonal deciduous forests of the north-east and central-west, and cerrados
and cerradões. The last two main groups contain caatingas, carrascos, chaco forests and peripheral
chaco seasonal forests.
The dendrogram for families goes a step further in generating groups with a strong vegetation
formation character. The semi-arid formations, namely the chaco forests, caatingas and carrascos, form a clump that merges, at the subsequent level, with a group that includes tropical and
subtropical seasonal deciduous forests and mesotrophic cerradões. An oddity of this group is
the presence of a side subgroup containing low-altitude rain forests and semideciduous forests
of the north-east and Amazonian regions. Tropical seasonal semideciduous forests predominate
in another main group that also includes the cerrado and tropical rain forests of the east and
north-east. The following main group contains tropical and subtropical rain forests and subtropical araucaria rain forests, and the last main group contains subtropical seasonal forests of
the south.
The tree floras represented in the rain and seasonal forest checklists are similar in species
richness: 3009 and 2903 species, respectively. On the other hand, the number of rain forest
checklists, 191, is considerably smaller than that of seasonal forests, 285, therefore suggesting that
the species richness of the latter may actually be lower. In fact, the speciesarea curves of the two
vegetation formations (Figure 7.8) demonstrate that, at any number of areas, the mean cumulative
number of species is much higher in rain than in seasonal forests. The two formations also share
a high proportion of tree species, 1814 out of 4098, or 44.3%, but both also have a considerable
number of putative endemics, 1195 (29.2%) and 1089 (26.6%) for rain and seasonal forests,
respectively.
The Venn diagrams in Figure 7.9 show the relationship of the tree flora of rain and seasonal
forests in different geographical regions. The number of species of both formations is smaller in
the north-east and south and larger in the east and south-east. The non-Atlantic seasonal forests
have higher numbers of species in the central-west and lower in the southwest. Seasonal deciduous
forests of the north-east, despite sharing many species with regional rain and semideciduous forests,
have their own group of putative endemics. The proportions of species shared by Atlantic rain and
seasonal forests are very similar in all regions: 20.4% in the north-east, 21.0% in the east, 22.0%
in the south-east, and 18.8% in the south. The species proportions in rain and seasonal forests,
2987_C007.fm Page 165 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
165
3500
Rain forests
3000
Seasonal
forests
Number of species
2500
2000
1500
1000
Caatingas
500
0
0
50
100
150
Number of areas
200
250
300
FIGURE 7.8 Mean cumulative number of species in areas of rain forests and seasonally dry tropical forests
of eastern South America with increasing number of areas.
however, show opposing trends from the north-east to the south, and are, respectively, 31.8% and
47.8% in the north-east, 39.2% and 39.8% in the east, 48.3% and 29.7% in the south-east, and
58.5% and 22.7% in the south. Subtropical araucaria rain forests share high proportions of species
with both rain and seasonal forests in both the south and South-east regions. However, they also
contain their group of putative endemics, particularly in the south. The central-west seasonal forests
share 76.2% of their species with Atlantic tropical rain- and seasonal forests (north-east, east and
south-east), though 60.6% are present in both formations, 15.6% in seasonal forests only, and none
in rain forests only.
The seasonal forests of all six geographical regions contain 2903 species, of which only 40 are
registered in all regions, 81 in five regions, 257 in four, 414 in three, 630 in two, and 1481 in one.
These putative endemic species are in higher proportion in the floras of the east (619; 35.0%) and
north-east (241; 32.1%), followed by the south-west (68; 27.1%), central-west (310; 24.4%), south
(66; 15.7%) and south-east (177; 14.9%). The relationships among seasonal forests of adjacent
geographical regions are shown in the left-side Venn diagrams of Figure 7.10. The three regions
of Atlantic tropical seasonal forests share a small number of species, 273 out of 2062 (13.2%), but
adjacent regions share larger proportions: 24.8% between the north-east and east and 29.0% between
the east and south-east. Subtropical seasonal forests share a high number of species with the tropical
seasonal forests of the south-east: 76.3% and 50.2% for the south and south-west, respectively. The
latter also share 59.8% of their species with the tropical seasonal forests of the central-west and
67.7% with both the South-east and central-west.
Caatingas are considerably poorer in tree species than are rain- and seasonal forests (Figure 7.8).
The relationships between caatingas and adjacent vegetation formations are shown in the right-side
Venn diagrams of Figure 7.10. A high proportion of their 466 species is shared with adjacent
seasonal forests, 61.2%, but this also leaves 38.8% of putative endemics. The proportion of shared
species is higher with the north-east seasonal forests (49.4%) than with the central-west (39.3%).
The proportion shared with cerrados is much smaller, 17.6%, and most of this is also shared with
central-west seasonal forests. The number of species shared with chaco forests is very small, only
five. In fact, both semi-arid formations, caatingas and chaco forests, share more species with the
central-west seasonal forests than between themselves. The geographical range of the tree flora of
2987_C007.fm Page 166 Thursday, December 1, 2005 7:03 PM
166
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Rain and seasonal forests-Northeast:
Deciduous
(236)
Rain and seasonal forests-East:
77
12
65
82
834
179
Rain (501)
22 8
954
849
290
Rain (1778)
Semideciduous (665)
Rain and seasonal forests-Southeast:
Rain and seasonal forests-South:
Araucaria rain (214)
23
Semideciduous (1803)
Araucaria rain (577)
19
131
165
15 6
16
971
676
458
224
298
57
77
Rain (1826)
Rain (890)
Semideciduous (1146)
73
Deciduous +
semideciduous (431)
Tropical rain and seasonal forests:
303
198
1120
Deciduous +
semideciduous CW
(1272)
771
531
Rain NE/E/SE
(2802)
911
Semideciduous
NE/E/SE (2411)
FIGURE 7.9 Venn diagrams extracted from the checklists showing the number of tree species shared by rainand seasonal forests in different geographical regions of eastern South America.
those two formations is illustrated in the two flow diagrams of Figure 7.11. Lowland seasonal
deciduous forests and carrascos of the north-east are excluded because they have a strong floristic
identity with the caatingas and are not in the route between the caatinga and chaco domains.
Seasonal forests of the east, South-east and south are merged into the category Austro-Atlantic
seasonal forests. The proportion of endemic to non-endemic species is lower for caatingas than is
for the chaco, 46.1:53.9% and 66.1:33.9%, respectively. The five species that complete the crossing
2987_C007.fm Page 167 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
Seasonal forests NE × E × SE:
167
Seasonal forests and caatingas:
Northeast (754)
Caatingas (466)
308
169
181
4
55
98
132
273
792
856
229
280
295
236
East (1823)
Southeast (1146)
Seasonal forests SE × S × SW:
Seasonal-CW (1272)
Seasonal-NE (754)
Seasonal forests, caatingas and cerrados:
Southwest (251)
Caatingas (466)
110
117
267
33
784
15
93
70
775
12
310
87
236
Southeast (1146)
South (431)
Seasonal forests SE × CW × SW:
Southwest (251)
20
174
Seasonal-CW (1272) Cerrados (1272)
Seasonal forests, caatingas and chaco:
81
44
Caatingas (466)
106
278
183
1
453
Southeast (1146)
567
555
Central-West (1272)
1058
4
27
151
Seasonal-CW (1272) Chaco (183)
FIGURE 7.10 Venn diagrams extracted from the checklists showing the number of tree species shared by
seasonal forests in different geographical regions of eastern South America (left side), and by seasonal forests,
caatingas, cerrados and chaco forests (right side).
between the two formations are Celtis pubescens (Jacquin) Sargent, Ximenia americana L.,
Sideroxylon obtusifolium (Roem. and Schultz) T.D.Penn., Parkinsonia aculeata L. and Aporosella
chacoensis (Morong) Pax and Hoffmg. All five are also present in the peripheral chaco seasonal
forests (south-west), but the former three also cross both the central-west and Austro-Atlantic
2987_C007.fm Page 168 Thursday, December 1, 2005 7:03 PM
168
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
1
AustroAtlantic
Seasonal
forests
63
Caatingas
N = 466
Endemics:
215
143
44
CentralWestern
Seasonal
forests
1
3
AustroAtlantic
Seasonal
forests
CentralWestern
Seasonal
forests
Peripheral
Chaco
Seasonal
forests
46
Chaco
forests
3
6
1
1
CentralWestern
Seasonal
forests
9
54
Chaco
forests
N = 183
Endemics:
121
8
Peripheral
Chaco
Seasonal
forests
CentralWestern
Seasonal
forests
AustroAtlantic
Seasonal
forests
AustroAtlantic
Seasonal
forests
12
14
3
Caatingas
CentralWestern
Seasonal
forests
1
FIGURE 7.11 Geographical extension of caatinga species towards seasonal forests and chaco
forests (top), and of chaco forest species towards seasonal forests and the caatingas (bottom).
Encircled figures are the number of species shared by the vegetation formations connected by
arrows.
seasonal forests while P. aculeata skips these two formations and A. chacoensis is also present in
central-west seasonal forests. Despite this, a much larger number of caatinga species, 55, reach the
peripheral chaco seasonal forests without entering the chaco itself. Similarly, 42 chaco species
reach the Austro-Atlantic and/or central-west seasonal forests without entering the caatingas.
The most species-rich genera and families of each main vegetation formation are given in
Tables 7.3 to 6, and the most frequent species of the same formations are provided in the Appendix.
Some genera rank high in most main Atlantic seasonal forest formations, e.g. Eugenia, Myrcia,
Ocotea and Miconia (Table 7.3). Some trends can be observed with increasing altitude: the relative
importance decreases for some genera such as Eugenia (except in the north-east), Inga and Ficus,
and increases for others such as Miconia and Tibouchina (east and south-east), Ilex and Solanum.
Subtropical seasonal forests of the south are similar to tropical seasonal forests in their generic
profile, but the south-west subtropical seasonal forests, chaco forests and the caatingas have very
particular sets of species-richest genera (Table 7.4). Among the families, Fabaceae is top in all
vegetation formations except the subtropical seasonal forests where it switches places with
Myrtaceae (Tables 7.5 and 7.6). In all other formations, Myrtaceae ranks second, except in the
south-west subtropical seasonal forests (3rd), chaco forests (16th), and caatingas (5th). Other
families ranking high among tropical and subtropical seasonal forests (except in the south-west)
are Rubiaceae, Melastomataceae and Lauraceae. Families showing increasing importance at higher
Swartzia
Copaifera
Croton
Clusia
Licania
Eugenia
16
12
9
9
9
9
9
9
8
8
8
8
7
7
7
6
6
6
6
5
5
5
5
5
5
Caesalpinia
Bauhinia
Clusia
Pilosocereus
Capparis
Psychotria
Psidium
Inga
Cyathea
Aspidosperma
Solanum
Casearia
Helicteres
Byrsonima
Acacia
Maytenus
Zanthoxylum
Ouratea
Croton
Ocotea
Senna
Cordia
Myrcia
Erythroxylum
High Altitude
(N = 11)
S
542
North-East Region
Tabernaemontana
Zanthoxylum
Coccoloba
Guapira
Mimosa
Ocotea
Bauhinia
Tabebuia
Pouteria
Casearia
Ficus
Aspidosperma
Psidium
Miconia
Inga
Senna
Erythroxylum
Cordia
Myrcia
Eugenia
Low Altitude
(N = 13)
17
15
12
9
9
8
7
7
7
6
6
6
6
6
6
5
5
5
5
5
4
4
4
4
4
S
420
Nectandra
Ilex
Swartzia
Croton
Psidium
Campomanesia
Erythroxylum
Rudgea
Trichilia
Aspidosperma
Casearia
Maytenus
Cordia
Solanum
Psychotria
Guatteria
Pouteria
Tabebuia
Ficus
Inga
Machaerium
Myrcia
Miconia
Ocotea
Eugenia
Low altitude
(N = 29)
49
32
32
30
23
22
22
15
15
14
14
14
13
13
13
12
12
12
11
11
11
10
10
9
9
S
1317
Vochysia
Cupania
Campomanesia
Ficus
Byrsonima
Croton
Cordia
Ouratea
Nectandra
Cinnamomum
Psidium
Guatteria
Casearia
Psychotria
Maytenus
Inga
Solanum
Erythroxylum
Ilex
Machaerium
Tibouchina
Eugenia
Ocotea
Myrcia
Miconia
High altitude
(N = 26)
East Region
43
37
29
29
18
17
16
16
16
15
14
13
13
12
12
11
11
11
10
10
10
10
10
10
10
S
1193
Casearia
Mollinedia
Aspidosperma
Styrax
Cestrum
Pouteria
Myrsine
Bauhinia
Croton
Maytenus
Zanthoxylum
Psychotria
Senna
Piper
Tabebuia
Inga
Erythroxylum
Nectandra
Machaerium
Solanum
Ficus
Myrcia
Miconia
Ocotea
Eugenia
Low Altitude
(N = 47)
37
27
24
18
16
15
13
12
11
11
10
10
9
9
9
8
8
8
8
8
8
8
7
7
7
S
848
Symplocos
Gomidesia
Mollinedia
Croton
Cordia
Tabebuia
Aspidosperma
Psychotria
Myrsine
Trichilia
Leandra
Casearia
Inga
Erythroxylum
Tibouchina
Nectandra
Ilex
Piper
Machaerium
Ficus
Solanum
Myrcia
Eugenia
Ocotea
Miconia
High Altitude
(N = 35)
South-East Region
36
26
26
18
16
15
14
14
12
11
11
10
10
10
9
9
9
9
8
8
8
8
8
8
8
S
911
Capparis
Vochysia
Cupania
Dalbergia
Tabebuia
Zanthoxylum
Acacia
Senna
Maytenus
Cordia
Psidium
Trichilia
Nectandra
Casearia
Byrsonima
Erythroxylum
Ocotea
Bauhinia
Inga
Aspidosperma
Ficus
Machaerium
Myrcia
Eugenia
Miconia
Low Altitude
(N = 74)
31
26
23
20
18
16
15
14
14
13
12
12
11
11
11
10
10
10
10
10
8
8
8
8
7
S
1129
Calyptranthes
Styrax
Alibertia
Dalbergia
Licania
Tabebuia
Guapira
Psidium
Trichilia
Byrsonima
Vochysia
Campomanesia
Maytenus
Inga
Ilex
Symplocos
Casearia
Nectandra
Ficus
Ocotea
Aspidosperma
Eugenia
Machaerium
Myrcia
Miconia
High Altitude
(N = 23)
Central-West Region
TABLE 7.3
Genera With the Highest Number of Species (S) in the Tree Flora of Tropical Seasonal Forests of the South American Atlantic Forest
Domain Classified into Four Geographical Regions and Two Altitudinal Ranges (N = Number of Areas)
28
18
13
12
11
11
11
10
9
9
8
8
7
7
7
6
6
6
6
5
5
5
5
5
5
S
624
2987_C007.fm Page 169 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
169
2987_C007.fm Page 170 Thursday, December 1, 2005 7:03 PM
170
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
TABLE 7.4
Genera with the Highest Number of Species (S) in the Tree Flora of Subtropical
Seasonal Forests of the South American Atlantic Forest Domain, Chaco Forests,
and Caatingas (N = Number of Areas)
Subtropical Seasonal Forests
South Region
(N = 37)
S
542
Eugenia
Ocotea
Myrsine
Tabebuia
Erythroxylum
Ficus
Myrcia
Solanum
Ilex
Maytenus
Sebastiania
Machaerium
Trichilia
Myrciaria
Psychotria
Zanthoxylum
Cestrum
Schinus
Cordia
Lonchocarpus
Inga
Nectandra
Miconia
Gomidesia
Symplocos
18
11
8
7
7
7
7
7
5
5
5
5
5
5
5
5
5
4
4
4
4
4
4
4
4
Southwest Rregion
(N = 5)
Acacia
Schinus
Aspidosperma
Tabebuia
Zanthoxylum
Chloroleucon
Ficus
Eugenia
Schinopsis
Tecoma
Maytenus
Bauhinia
Mimosa
Ceiba
Luehea
Cedrela
Trichilia
Myrsine
Myracrodruon
Ilex
Ruprechtia
Prosopis
Ziziphus
Capparis
Carica
S
420
Chaco Forests
(N = 39)
S
1193
Caatingas
(N = 6)
S
1317
9
7
6
5
5
4
4
4
3
3
3
3
3
3
3
3
3
3
2
2
2
2
2
2
2
Prosopis
Acacia
Lycium
Echinopsis
Opuntia
Jatropha
Senna
Capparis
Bougainvillea
Bulnesia
Schinopsis
Cereus
Harrisia
Maytenus
Caesalpinia
Mimosa
Aloysia
Myrcianthes
Ruprechtia
Condalia
Zanthoxylum
Quiabentia
Ceiba
Aspidosperma
Berberis
19
9
8
7
7
7
6
5
5
5
4
4
4
4
4
4
4
3
3
3
3
2
2
2
2
Croton
Mimosa
Senna
Erythroxylum
Bauhinia
Manihot
Eugenia
Aspidosperma
Cordia
Pilosocereus
Acacia
Helicteres
Tabebuia
Maytenus
Caesalpinia
Psidium
Zanthoxylum
Capparis
Pereskia
Cnidoscolus
Hymenaea
Chloroleucon
Guapira
Ruprechtia
Facheiroa
14
13
11
10
9
8
8
7
7
7
7
7
6
6
6
6
5
4
4
4
4
4
4
3
3
altitudes are Asteraceae, Melastomataceae (except in the north-east) and Lauraceae (though
unchanged in the east and south-east). Euphorbiaceae are particularly important in most formations
but rank higher in the north-east and central-west low-latitude seasonal forests, as well as in the
caatingas and chaco forests, which are also distinguished by the high ranking of Cactaceae.
7.4 DISCUSSION
An overall pattern emerging from the floristic analyses of the vegetation formations of eastern South
America was the strong influence of distance on tree species distribution. This influence only receded
and allowed clear climate-related patterns to be discerned when either the geographical range considered was restricted or data were treated at generic and familial levels. Likewise, geographically
restricted analyses of Atlantic forest sections, such as those performed for south-east Brazil by
Oliveira-Filho and Fontes (2000) and north-east Brazil by Ferraz et al. (2004), could clearly detect
species patterns primarily related to the climate. However, analyses performed for wider geographical
ranges, such as the Amazon, could best detect patterns related to climate and vegetation formations when
dealing with genera and families rather than species (ter Steege et al., 2000; Oliveira and Nelson, 2001).
S
542
113
48
28
19
18
17
16
15
14
13
11
11
11
10
10
10
9
9
9
9
9
8
7
7
6
Low Altitude
(N = 13)
Fab
Myrt
Rubi
Euphorbi
Apocyn
Malv
Sapot
Annon
Mor
Clusi
Bignoni
Chrysobalan
Salic
Boragin
Melastomat
Sapind
Anacardi
Combret
Erythroxyl
Laur
Rut
Arec
Nyctagin
Polygon
Lecythid
Fab
Myrt
Rubi
Euphorbi
Erythroxyl
Rut
Malv
Laur
Solan
Boragin
Malpighi
Anacardi
Cact
Clusi
Salic
Apocyn
Aster
Bignoni
Ochn
Celastr
Chrysobalan
Melastomat
Mor
Sapind
Cyathe
High Altitude
(N = 11)
North-East Region
81
44
22
17
15
12
11
10
10
9
9
8
8
8
8
7
7
7
7
6
6
6
6
6
5
S
420
Fab
Myrt
Rubi
Laur
Melastomat
Annon
Euphorbi
Mor
Sapot
Solan
Aster
Bignoni
Malv
Sapind
Arec
Clusi
Rut
Apocyn
Celastr
Chrysobalan
Meli
Salic
Anacardi
Nyctagin
Boragin
Low Altitude
(N = 29)
192
140
83
68
48
44
44
35
31
29
27
27
26
25
24
24
24
23
19
19
19
19
16
14
13
S
1317
Fab
Myrt
Melastomat
Laur
Rubi
Aster
Euphorbi
Annon
Solan
Clusi
Sapind
Malv
Rut
Apocyn
Chrysobalan
Mor
Bignoni
Salic
Vochysi
Celastr
Aquifoli
Erythroxyl
Sapot
Malpighi
Meli
High Altitude
(N = 26)
East region
151
137
73
65
58
51
38
32
27
24
24
21
20
19
19
19
18
18
18
17
16
16
16
14
13
S
1193
Fab
Myrt
Rubi
Laur
Melastomat
Euphorbi
Solan
Rut
Mor
Annon
Bignoni
Malv
Salic
Sapind
Aster
Myrsin
Meli
Anacardi
Celastr
Erythroxyl
Sapot
Apocyn
Lami
Piper
Vochysi
Low Altitude
(N = 47)
111
101
55
49
34
31
29
24
22
18
17
16
16
16
15
14
12
11
11
11
11
10
10
10
10
S
848
Fab
Myrt
Melastomat
Laur
Rubi
Aster
Euphorbi
Solan
Mor
Malv
Annon
Rut
Salic
Bignoni
Vochysi
Meli
Piper
Sapind
Myrsin
Apocyn
Aquifoli
Celastr
Clusi
Erythroxyl
Cyathe
High Altitude
(N = 35)
South-east region
108
96
59
52
44
33
32
27
21
20
19
19
17
15
15
14
14
14
13
12
12
12
10
10
9
S
911
Fab
Myrt
Rubi
Euphorbi
Melastomat
Malv
Annon
Laur
Mor
Sapind
Rut
Apocyn
Salic
Nyctagin
Meli
Chrysobalan
Arec
Bignoni
Vochysi
Celastr
Aster
Combret
Malpighi
Sapot
Boragin
Low altitude
(N = 74)
210
95
58
46
39
36
32
32
27
27
26
25
21
18
18
17
17
16
16
16
16
15
15
14
13
S
1129
Fab
Myrt
Rubi
Melastomat
Laur
Annon
Malv
Mor
Vochysi
Salic
Apocyn
Arec
Bignoni
Celastr
Clusi
Euphorbi
Meli
Chrysobalan
Sapind
Aster
Myrsin
Nyctagin
Anacardi
Symploc
Aquifoli
High altitude
(N = 23)
Central-west region
93
62
37
34
29
20
17
16
15
14
13
13
12
12
12
12
12
11
11
10
10
10
9
9
8
S
624
TABLE 7.5
Families with the Highest Number of Species (S) in the Tree Flora of Tropical Seasonal Forests of the South American Atlantic
Forest Domain Classified into Four Geographical Regions and Two Altitudinal Ranges (Suffix ‘-aceae’ Omitted from All Families;
N = Number of Areas)
2987_C007.fm Page 171 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
171
2987_C007.fm Page 172 Thursday, December 1, 2005 7:03 PM
172
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
TABLE 7.6
Families with the Highest Number of Species (S) in the Tree Flora of Subtropical
Seasonal Forests of the South American Atlantic Forest Domain, Chaco Forests
and Caatingas (Suffix ‘-aceae’ Omitted from All Families; N = Number of Areas)
Subtropical seasonal forests
South Region
(N = 37)
Myrt
Fab
Laur
Rubi
Solan
Euphorbi
Bignoni
Meli
Mor
Aster
Salic
Rut
Myrsin
Anacardi
Erythroxyl
Malv
Melastomat
Sapind
Urtic
Celastr
Sapot
Apocyn
Aquifoli
Arec
Boragin
S
542
60
49
21
18
18
17
11
11
11
10
10
9
8
7
7
7
7
7
7
6
6
5
5
5
5
South-west Region
(N = 5)
Fab
Anacardi
Myrt
Malv
Sapind
Bignoni
Apocyn
Euphorbi
Rut
Meli
Mor
Salic
Arec
Celastr
Nyctagin
Sapot
Aster
Cannab
Laur
Polygon
Rubi
Brassic
Rhamn
Myrsin
Phytolacc
S
420
Chaco Forests
(N = 39)
S
1193
Caatingas
(N = 6)
S
1317
55
14
14
13
12
10
8
8
7
6
6
6
5
5
5
5
4
4
4
4
4
3
3
3
3
Fab
Cact
Solan
Euphorbi
Zygophyll
Anacardi
Nyctagin
Rhamn
Brassic
Celastr
Malv
Verben
Arec
Bignoni
Cannab
Myrt
Rut
Sapind
Sapot
Apocyn
Malpighi
Mor
Polygon
Santal
Aster
63
29
13
11
9
7
7
6
5
5
5
5
4
4
4
4
4
4
4
3
3
3
3
2
2
Fab
Euphorbi
Cact
Malv
Myrt
Apocyn
Bignoni
Erythroxyl
Boragin
Rubi
Rut
Celastr
Combret
Annon
Nyctagin
Polygon
Sapind
Arec
Brassic
Anacardi
Rhamn
Salic
Solan
Aster
Malpighi
106
36
24
19
16
11
10
10
9
9
8
7
7
6
6
6
6
5
5
4
4
4
4
3
3
The geographical proximity among different vegetation formations within the same region and evolution through adaptive radiation into adjacent habitats could explain much of the strong effect of
distance found in species patterns throughout the geographical range analysed here. On the other
hand, the patterns related to climate and vegetation formations found for genera and families strongly
suggest that climatic variables, particularly temperature and water availability, have had a long
influence on the evolution and speciation of tree taxa in eastern South America. This is not a surprise
since water and temperature are the chief factors determining the distribution of most world’s vegetation
formations, and the history of vegetation and climate of that part of the world during the Quaternary
shows dramatic shifts in both temperature and rainfall regime (Salgado-Labouriau et al., 1997;
Behling, 1998; Ledru et al., 1998, Oliveira et al., 1999).
One important result of the above-mentioned geographical pattern is that there was greater
similarity in species composition between Atlantic rain and seasonal forests of the same region
than between either seasonal or rain forests of disjunct regions, although this holds true only when
east and south-east are merged. In the same region the tree flora of seasonal forests is much less
diverse than that of the rain forests, and is probably composed of species able to cope with relatively
longer dry seasons. Tree species diversity in tropical forests is often correlated with water
2987_C007.fm Page 173 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
173
consumption and energy uptake, resources that are partitioned among species and limit their number
in forest communities (Hugget, 1995). Water shortage probably plays the chief role in reducing
species richness of seasonal forests compared to rain forests, and even more so of semi-arid
formations such as chaco forests, caatingas and carrascos. Moreover, the structure of seasonal
forests is also less complex, therefore favouring a reduced number of understory species compared
to rain forests (Gentry and Emmons, 1987).
The intimate relationship between the two floras within each geographical region supports the
wider definition of Atlantic forests to include both rain and seasonal forests as physiognomic and
floristic expressions of a single great vegetation domain (Oliveira-Filho and Fontes, 2000; Galindo-Leal
and Câmara, 2003). For all regions but the north-east, there was a greater floristic similarity at all
three taxonomic levels between Atlantic rain and seasonal forests than between either of these and
Amazonian rain forests. The exceptions were the north-east rain and seasonal semideciduous forests,
both closer to Amazonian rain forests though only at the generic and familial levels. As the coastal
north-east is climatically and geographically closer to the Amazon, a stronger past link could have
existed through the so-called north-east bridge (Bigarella et al. 1975; Mori et al., 1981; AndradeLima, 1982; Cavalcanti and Tabarelli, 2004). Nevertheless, as shown by the present results, this
alleged link also includes rain and seasonal forests and, for that reason, there is little floristic ground
for viewing Atlantic rain forests as being closer to Amazonian rain forests than to their adjacent
seasonal forests.
In all four Atlantic regions, seasonal forests and their rain forest neighbours share a similar
proportion of the total species count (c.20%) and are both poorer in species in the north-east and
south and much richer in the east and south-east. Seasonal forests of the central-west were also
comparatively rich. An inspection of the distribution map (Figure 7.1) helps us understand this. Of
all regions, the north-east has the smallest forest area and also lacks the highly rugged relief of
other regions, the latter being a feature that may boost species richness through increased environmental heterogeneity. In addition, the region may have lost much of its primitive species richness
because it was the first to go through mass deforestation, beginning in the sixteenth century. It is
the least known, and most threatened and reduced of all Atlantic forests, now covering only 3.76%
of its original area (Silva and Tabarelli, 2000, 2001). Towards the south, Atlantic seasonal forests
expand increasingly more into the continental interior until reaching Mato Grosso do Sul and
eastern Paraguay, so that they cover a wide area with pronounced variation in relief and climate
(Oliveira-Filho and Fontes, 2000). In addition, seasonal forests also spread towards the west into
the whole of the cerrado domain as galleries and forest patches that are found as far as in the
Bolivian chiquitanía (see Kileen et al., Chapter X). The high environmental heterogeneity of this
large area, combined with the complex contact with the cerrado, certainly explains the high species
richness of the east, south-east and central-west tropical seasonal forests. The comparatively lower
speciesrichness of the subtropical seasonal forests of the south may also be explained by their
comparatively smaller area and modest relief, but these forests are also at the southernmost range
of Atlantic forests where extreme low temperatures in winter coupled with frosts may have already
selected the small proportion of tree species able to cope with this climatic harshness (Rambo,
1980; Leite, 2002; Jurinitz and Jarenkow, 2003).
Surprisingly the proportion of seasonal forest species shared with rainforests remains more or
less constant throughout the geographical range, despite the increase in species numbers of rainforest from north-east and north to south-east and south. This is brought about by an increase in
the percentage of seasonal forest species occurring in rainforest from 51% and 52.9% in the northeast and east, respectively, to 74% and 83% in the south-east and south, thus counterbalancing the
increase in rainforest endemics and maintaining the proportion. Unlike the situation in the other
regions, the number of species in the north-east seasonal forests actually surpasses that of rain
forests. The main contrast between the two pairs of northern and southern regions is that the
former (north-east and east) have warmer temperatures and a much more pronounced variation
in rainfall totals and seasonality than the latter (south-east and south) since they are adjacent to
2987_C007.fm Page 174 Thursday, December 1, 2005 7:03 PM
174
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
the caatinga domain. This is probably why north-east seasonal forests are the only ones to show a
clear distinction between semideciduous and deciduous formations. The wider rainfall gradient is
correlated with a relatively rapid transition from rain- to semideciduous and deciduous forests, and
from those to caatingas, and thus because of ecotonal effects increasing the speciesrichness of
seasonal forests relative to their ‘purer’ rain forest partners. This explains, for example, why
Cactaceae and Euphorbiaceae are so important in the flora of montane semideciduous forests, the
so-called brejo forests, which occur as hinterland forest islands on mountains surrounded by lowland
caatingas, and inevitably share a number of species with the latter (Rodal, 2002; Pôrto et al., 2004).
In addition to this, the assemblage of north-east seasonal forests includes those influenced by other
neighbouring formations, such as the coastal sandy restingas (e.g. at Fernando de Noronha and
Natal) and the cerrado (e.g. at Araripe, Campo Maior and Sete Cidades) that may also boost their
species richness (Farias and Castro, 2004). Similar effects may have occurred in the eastern region
which also combines floristic interactions with both caatingas and cerrados, in addition to the effects
of the rugged relief of the Espinhaço mountain range and the Chapada Diamantina (Guedes and
Orge, 1998; Zappi et al., 2003).
The central-west seasonal forests also have strong floristic links with both the cerrados and
Atlantic forests and share a considerable number of species. In fact, one could extract a continuum
in tree species distribution determined by rainfall seasonality starting at the east and south-east
Atlantic rain and seasonal forests, and extending towards the central-west to reach its seasonal
forests and, lastly, the cerradões and cerrados, as already suggested by Leitão-Filho (1987). However,
this is an oversimplified view of the floristic gradient because it is now known that, under seasonal
climates, more important factors are involved in determining the forest-cerrado transition, and fire
frequency and soil fertility and moisture play the chief role here (Oliveira-Filho and Ratter, 2002). As
a result, it is not uncommon to find in the central-west two or more of those formations on a single
slope (Furley and Ratter, 1988; Furley et al., 1988; Ratter et al., 1978). The complex mosaic of
vegetation formations of the region and the species interchange among them probably explain why
the analyses failed to discriminate deciduous from semideciduous forests floristically. Although the
two formations do form a continuum it is usually easy to tell at least their extremes apart in the
field on the basis of physiognomy and floristic composition (Oliveira-Filho and Ratter, 2002). For
that reason, particular attention should be paid to such nuances in the preparation of checklists for
the region.
Towards the south and south-east, the declining temperatures and related vapor pressure curtail
the water deficit gradient and this probably favors a stronger floristic relationship between rain and
seasonal forests expressed by the much higher proportions of shared species. Extremes of low
temperatures may be important determinants of tree species distribution. Occasional frosts have
been mentioned by Oliveira-Filho et al. (1994) as an important factor limiting species distribution
both in relation to higher elevations and latitudes in south and south-east Brazil. Resistance to
frosts was suggested as a key factor determining the special nature of the chaco flora, together with
their saline to alkaline soils (Pennington et al., 2000). The influence of latitude and altitude on
climate, however, is far more complex than simply that of temperature and frosts. Increasing latitude
also means increasing year-round variation of the daily sunlight period. Rising elevation also
decreases atmospheric pressure, increases solar radiation, accelerates windmovement, promotes
greater cloudiness and boosts rainfall (Jones, 1992). For tropical forests, rainfall seasonality is
apparently more important than annual rainfall in determining presence of rain or seasonal forests,
and the occurrence of at least a 30-day dry season produces effects which can clearly be shown
on a vegetation map (IBGE, 1993). For subtropical forests, however, temperature range prevails
over rainfall seasonality in separating rain and seasonal forests, probably because the contrast
of low winter and high summer temperatures plays an additional role in forest deciduousness
(Holdridge et al., 1971). Low temperatures alone, without the strong annual oscillation, are not
associated with the presence of subtropical seasonal forests, and other formations appear, in particular
araucaria rain forests, in the hinterland highlands, and upper montane rain forests (cloud forests) on
2987_C007.fm Page 175 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
175
the mountain ridges near the coast (Roderjan et al., 2002). Semi-arid formations such as chaco
forests are found only where very strong rainfall seasonality occurs in subtropical climates, but
under these conditions forests also give way to open grasslands (campos or pampas) in many areas
of the south, and this is probably linked to the past history of fire and grazing (Behling, 1995,
1997; Quadros and Pillar, 2002).
Tree species composition of seasonal forests is highly influenced by altitude and associated
temperatures, a well-known fact for mountain vegetation worldwide (Hugget, 1995). Because most
mountain ranges and plateaus in our area of study are concentrated in the east and the lowlands of
the Paraguay river basin lie in the west, the seasonal forest gradient related to decreasing altitude
and increasing temperatures is highly coincident with increasing distance from the ocean. For this
reason, one might speculate that this gradient was another primarily related to distance. Nevertheless, altitude-related gradients have already been detected for Atlantic rain and seasonal forests at
more regional scales by Oliveira-Filho and Fontes (2000) and Ferraz et al. (2004) in south-east and
north-east Brazil, respectively, and by Salis et al. 1995, Torres et al. 1997 and Scudeller et al. (2001)
in the state of São Paulo. Moreover, some floristic patterns found with increasing altitude also
coincided with those cited by the above authors and by Gentry (1995) for Andean and Central
American forests. Among these, are the increasing contribution of Melastomataceae to the tree
flora, particularly Miconia and Tibouchina, Solanaceae (Solanum), Lauraceae (Ocotea and Nectandra), Aquifoliaceae (Ilex) and Asteraceae, and the decrease of Eugenia and Ficus with increasing
altitude. A detailed treatment of genera and species diagnostic of montane Atlantic forests is given
by Oliveira-Filho and Fontes (2000).
It is now accepted that the caatinga domain represents the largest, most isolated and speciesrich nucleus of the SDTF and that its flora is made up of a blend of endemic and wide-range species
(Giulietti et al., 2002; Prado, 2003). The strongest internal floristic dichotomy of the semi-arid
vegetation of the caatinga is that linked to soils derived from either crystalline base rock or sandy
deposits (Araújo et al., 1998; Rodal and Sampaio, 2002; Rocha et al., 2004). The present analyses
largely support these findings, which are considered by Queiroz in Chapter X. He hypothesizes
that the vast proportion of caatinga non-endemics results from post-Tertiary migration of widerange SDTF species into the region as a result of the progressive retreat of sandy deposits and
exposure of the crystalline bedrock. In fact, most non-endemic caatinga species are found throughout the Austro-Atlantic and central-west seasonal forests and many reach the peripheral chaco
seasonal forests without entering the chaco itself, as already emphasized by Prado (1991) and Prado
and Gibbs (1993) to demonstrate the strong differentiation of the chaco and caatinga floras.
Interestingly, our analyses also suggest that a similar pattern may also occur in non-endemic chaco
species that are also found in Austro-Atlantic and/or central-west seasonal forests but do not enter
the caatingas. Despite this similarity there is also an important difference in that most non-endemic
chaco species show a more limited distribution outside the chaco domain and do not reach as far
as the caatinga periphery. Thus, chaco and caatinga are well-defined floristic nuclei with very weak
relationships between their floras at both specific and generic levels. Only at the family level do
the two floras show a stronger link, thus suggesting that if a common proto-flora did exist it must
have been in the very remote past. Both floras also show floristic connections with the Atlantic and
central-west seasonal forests but this is probably mainly the result of both species interchange in
transitional areas and expansion of wide-range SDTF species.
7.5 CONCLUSION
We propose here that one can best describe Atlantic seasonal forests as a section of a complex
floristic gradient extending from evergreen forests to semideciduous and deciduous forests (the
SDTF section), and then running in a partial blending of floras to open formations, such as cerrados
and campos or, alternatively, to caatingas and chaco forests. This gradient is chiefly related to
decreasing water availability through either increasing rainfall seasonality and/or decreasing soil
2987_C007.fm Page 176 Thursday, December 1, 2005 7:03 PM
176
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
moisture content, but there is also a strong interference of temperature gradients along the latitudinal
and altitudinal ranges, and of variations in soil fertility and fire frequency. The flora of the Atlantic
seasonal forests occurs mostly in the section of the tropical gradient corresponding to annual periods
of water shortage between 30 and 160 days, and in the section of the subtropical range where
periods of water shortage are below 30 days but year-round monthly temperature oscillation is
above 10°C. Increasing periods of water shortage, soil fertility and temperature range normally
lead from semideciduous to deciduous forests and then to the semi-arid formations, either caatingas
(tropical) or chaco forests (subtropical), while increasing fire frequency and decreasing soil fertility
frequently lead from seasonal forests to either cerrados (tropical) or southern campos (subtropical).
For this reason we suggest here that the definition of SDTF should be reshaped to include both
cerrados and the chaco.
In conclusion, we believe that the best view of the SDTF vegetation of eastern South America
is that of three floristic nuclei: caatinga, chaco and Atlantic forest (sensu latissimo). Only the latter,
however, should be linked consistently to the residual Pleistocenic dry seasonal (RPDS) flora.
Caatinga and chaco form the extremes of floristic dissimilarity among the three SDTF nuclei, also
corresponding to the warm-dry and warm-cool climatic extremes, respectively. In contrast, the
Atlantic SDTF nucleus is poor in endemic species and is actually a floristic bridge connecting the
two drier nuclei to rain forests. Additionally, there is little evidence to describe the Atlantic nucleus
flora as a clearly distinct species assemblage, as are those of the caatinga and chaco nuclei, because
of the striking variation in species composition found throughout the vast geographical extent of
Atlantic seasonal forests. Nevertheless, there is a group of wide-range species that is found in most
regions of the Atlantic nucleus, some of which are also part of the species blend of the caatinga
and chaco floras, though involving the latter to a much lesser extent. We believe that, at least in
eastern South America, it is precisely this small fraction of the Atlantic nucleus flora that should
be identified with the RPDS vegetation. To clarify the past history of neotropical SDTF, we propose
that the focus should now be shifted to the investigation of the distribution patterns of those species
and the past history of their populations in different locations of their geographical range.
ACKNOWLEDGEMENTS
The first author thanks the CNPq and the Royal Society of Edinburgh for the financial support to
the present study and the Royal Botanic Garden Edinburgh for warmly hosting him once more.
We were helped during the taxonomic revision of the database by the following: Marcos Sobral
(Myrtaceae) and João Renato Stehmann (Solanaceae), both from the Federal University of
Minas Gerais; Haroldo Lima (Fabaceae), Alexandre Quinet (Lauraceae), José Fernando Baumgratz
(Melastomataceae) and Angela Studart da Fonseca Vaz (Bauhinia) from the Rio de Janeiro Botanic
Garden; Maria Célia Vianna (Vochysia) from the Alberto Castellanos Herbarium; José Rubens Pirani
(Simaroubaceae, Picramniaceae and Rutaceae) and Pedro Fiaschi (Araliaceae) from the University
of São Paulo; Maria Lúcia Kawazaki (Myrtaceae) and Inês Cordeiro (Euphorbiaceae) from the São
Paulo Botanic Institute; Washington Marcondes-Ferreira (Aspidosperma) from the State University
of Campinas; Germano Guarim Neto (Cupania) from the Federal University of Mato Grosso; and
Toby Pennington and Maureen Warwick (Fabaceae) from the Royal Botanic Garden Edinburgh.
We also thank Toby Pennington, James Ratter and an anonymous reviewer for their critical and
constructive reading of the first draft.
REFERENCES
Andrade-Lima, D., Present-day forest refuges in northeastern Brazil, in Biological Diversification in the
Tropics, Prance, G.T., Ed., Plenum Press, New York, 1982, 220.
APG, An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering
plants: APG II, Bot. J. Linn. Soc., 141, 399, 2003.
2987_C007.fm Page 177 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
177
Araújo, F.S. et al., Composição florística da vegetação de carrasco, Novo Oriente, CE, Rev. Brasil. Bot., 21,
105, 1998.
Behling, H., Investigations into the Late Pleistocene and Holocene history of vegetation and climate in Santa
Catarina (S. Brazil), Veget. Hist. Archaeo., 4, 127, 1995.
Behling, H., Late Quaternary vegetation, climate and fire history in the Araucaria forest and campos region
florm Serra Campos Gerais (Paraná) S. Brazil., Rev. Palaeobot. Palyn. 97, 109, 1997.
Behling, H., 1998. Late Quaternary vegetational and climatic changes in Brazil, Rev. Palaeobot. Palyn., 99,
143, 1998.
Bigarella, J.J., Andrade-Lima, D., and Riehs, P.J., Considerações a respeito das mudanças paleoambientais na
distribuição de algumas espécies vegetais e animais no Brasil, Anais Acad. Brasil. Ciênc., 47, 411,
1975.
ter Braak, C.J.F., Ordination, in Data Analysis in Community and Landscape Ecology, Jongman, R.H.G., ter
Braak, C.J.F. and van Tongeren, O.F.R., Eds., Cambridge University Press, Cambridge, 1995, 91.
Bridgewater, S. et al., A preliminary floristic and phytogeographic analysis of the woody flora of seasonally
dry forests in northern Peru, Candollea, 58, 129, 2003.
Cavalcanti, D. and Tabarelli, M. Distribuição das plantas Amazônico-Nordestinas no Centro de Endemismo
Pernambuco: brejos de altitude vs. florestas de terras baixas, in Brejos de Altitude em Pernambuco e
Paraíba, História Natural, Ecologia e Conservação, Pôrto, K.C., Cabral, J.J.P., and Tabarelli, M.,
Eds. Ministério do Meio Ambiente, Brasília, 2004, 285.
DNMet, Normais climatológicas (1961–1990). Departamento Nacional de Meteorologia, Ministério da Agricultura, Brasília, 1992.
Farias, R.R.S. and Castro, A.A.J.F., Fitossociologia da vegetação do Complexo de Campo Maior, Campo
Maior, PI, Brasil, Acta Bot. Brasil., 18, 949, 2004.
Ferraz, E.M.N., Araújo, E.L., and Silva, S.I., Floristic similarities between lowland and montane areas of
Atlantic Coastal Forest in Northeastern Brazil, Plant Ecol. 174, 59, 2004.
Furley, P.A. and Ratter, J.A., Soil resources and plant communities of the central Brazilian cerrado and their
development, J. Biogeogr., 15, 97, 1988.
Furley, P.A., Ratter, J.A., and Gifford, D.R. Observations on the vegetation of eastern Mato Grosso, Brazil.
III. The woody vegetation and soils of the Morro de Fumaça, Torixoréu. Proc. Roy. Soc. London B,
235, 259, 1988.
Galindo-Leal, C. and Câmara, I.G., Atlantic Forest hotspot status: an overview, in The Atlantic Forest of South
America, Galindo-Leal, C. and Câmara, I.G., Eds., Center for Applied Biodiversity Science,
Washington, 2003, 3.
Gentry, A.H., Patterns of diversity and floristic composition, in Neotropical Montane Forests, Biodiversity and
Conservation of Neotropical Montane Forests, Neotropical Montane Forest Biodiversity and Conservation Symposium, 1, Churchill, S.P. et al., Eds., The New York Botanical Garden, New York, 1995, 103.
Gentry, A.H. and Emmons, L.H., Geographical variation in fertility, phenology and composition of the
understory of Neotropical forests, Biotropica, 19, 216, 1987.
Giulietti, A.M. et al., Espécies endêmicas da Caatinga, in Vegetação e Flora da Caatinga, Sampaio, E.V.S.B.
et al., Eds., Associação Plantas do Nordeste, Recife, 2002, 103.
Graham, A. and Dilcher, D., The Cenozoic record of tropical dry forest in northern Latin America and the
southern United States, in Seasonally Dry Tropical Forests, Mooney, H.A., Bullock, S.H., and Medina,
E., Eds., Cambridge University Press, Cambridge, 1995, 124.
Guedes, M.L.S. and Orge, M.D.R. Checklist das espécies vasculares de Morro do Pai Inácio (Palmeiras) e
Serra da Chapadinha (Lençóis). Chapada Diamantina, Bahia. Instituto de Biologia da Universidade
Federal da Bahia, Salvador, 1998.
Hill, M.O. and Gauch, H.G. Detrended correspondence analysis: an improved ordination technique, Vegetatio,
42, 47, 1980.
Holdridge, L.R. et al., Forest Environments in Tropical Life Zones,A Pilot Study, Pergamon, Oxford, 1971.
Hugget, R.J., Geoecology, An EvolutionaryAapproach, Routledge, London, 1995.
IBGE, Mapa de Vegetação do Brasil. Instituto Brasileiro de Geografia e Estatística, Rio de Janeiro, 1993.
Jones, H.G., Plants and Microclimate: A Quantitative Approach to Environmental Plant Physiology, 2,
Cambridge University Press, Cambridge, 1992.
Jurinitz, C.F. and Jarenkow, J.A., Estrutura do componente arbóreo de uma floresta estacional na Serra do
Sudeste, Rio Grande do Sul, Brasil. Rev. Brasil. Bot., 26, 475, 2003.
2987_C007.fm Page 178 Thursday, December 1, 2005 7:03 PM
178
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Kent, M. and Coker, P., Vegetation Description andAanalysis, APpracticalAapproach, Belhaven Press, London,
1992.
Ledru, M.P., Salgado-Labouriau, M.L., and Lorscheiter. M.L., Vegetation dynamics in southern and central
Brazil during the last 10,000 yr.B.P., Rev. Palaeobot. Palyn. 99, 131, 1998.
Leitão-Filho, H.F., Considerações sobre a florística de florestas tropicais e sub-tropicais do Brasil. Rev. IPEF
35, 41, 1987.
Leite, P.F., Contribuição ao conhecimento fitoecológico do Sul do Brasil, Ciênc. Amb., 24, 51, 2002.
Lewis, J.P., Pire, E.F., and Vesprini, J.L., The mixed dense forest of the Southern Chaco. Contribution to the
study of flora and vegetation of the Chaco. VIII. Candollea, 49, 159, 1994.
Linares-Palomino, R., Pennington, R.T., and Bridgewater, S., The phytogeography of the seasonally dry
tropical forests in Equatorial South America, Candollea, 58, 473, 2003.
McCune, B. and Mefford, M.J., PC-ORD version 4. 0., Multivariate Analysis of Ecological Data, Users Guide.
MjM Software Design, Glaneden Beach, 1999.
Mooney, H.A., Bullock, S.H., and Medina, E., Introduction, in Seasonally Dry TropicalFforests, Mooney,
H.A., Bullock, S.H., and Medina, E., Eds., Cambridge University Press, Cambridge, 1995, 1.
Mori, S.A., Boom, B.M., and Prance, G.T., Distribution patterns and conservation of eastern Brazilian coastal
forest tree species, Brittonia, 33, 233, 1981.
Murphy, P.G. and Lugo, A.E., Dry forests of Central America and the Caribbean, in Seasonally Dry Tropical
Forests, Mooney, H.A., Bullock, S.H., and Medina, E., Eds., Cambridge University Press, Cambridge,
1995, 146.
Oliveira, A.A. and Nelson, B.W., Floristic relationships of terra firme forests in the Brazilian Amazon, Forest
Ecol. Manag., 146, 169, 2001.
Oliveira, P.E., Barreto, A.M.F., and Suguio, K., Late Pleistocene/Holocene climatic and vegetational history
of the Brazilian Caatinga: the fossil dunes of the middle São Francisco river, Palaeogeogr. Palaeoclimat. Palaeocecol., 152, 319, 1999.
Oliveira-Filho, A.T. and Fontes M.A.L., Patterns of floristic differentiation among Atlantic forests in southeastern Brazil, and the influence of climate, Biotropica, 32, 793, 2000.
Oliveira-Filho, A.T. and Ratter, J.A., Database: Woody Flora of 106 Forest Areas of Eastern Tropical South
America, Royal Botanic Garden Edinburgh, Edinburgh, 1994.
Oliveira-Filho, A.T. and Ratter, J.A., A study of the origin of central Brazilian forests by the analysis of plant
species distribution patterns, Edinburgh J. Bot., 52, 141, 1995.
Oliveira-Filho, A.T. and Ratter, J.A., Vegetation physiognomies and woody flora of the Cerrado Biome, in
The Cerrados of Brazil: Ecology and Natural History of a Neotropical Savanna, Oliveira, P.S. and
Marquis, R.J., Eds., Columbia University Press, New York, 2002, 91.
Oliveira-Filho, A.T. et al., Comparison of the woody flora and soils of six areas of montane semi-deciduous
forest in southern Minas Gerais, Brazil, Edinburgh J. Bot. 51, 355, 1994.
Pennington, R.T., Prado, D.E., and Pendry, C.A., Neotropical seasonally dry forests and Quaternary vegetation
changes, J. Biogeogr., 27, 261, 2000.
Pennington, R.T. et al., Historical climate change and speciation: neotropical seasonally dry forest plants show
patterns of both Tertiary and Quaternary diversification, J. Biogeogr., 359, 515, 2004.
Pôrto, K.C., Cabral, J.J.P., and Tabarelli, M., Brejos de Altitude em Pernambuco e Paraíba, História Natural,
Ecologia e Conservação, Ministério do Meio Ambiente, Brasília, 2004.
Prado, D.E., A critical evaluation of the floristic links between Chaco and Caatingas vegetation in South
America, PhD thesis, University of St. Andrews, St. Andrews, 1991.
Prado, D.E., Seasonally dry forests of tropical South America: from forgotten ecosystems to a new phytogeographic unit. Edinburgh J. Bot., 57, 437, 2000.
Prado, D.E., As Caatingas do Brasil, in Ecologia e Conservação da Caatinga, Leal, I.R., Tabarelli, M., and
Silva, J.M.C., Eds., Editora Universidade Federal de Pernambuco, Recife, 2003, 3.
Prado, D.E. and Gibbs, P.E., Patterns of species distribution in the dry seasonal forests of South America.
Ann. Mo. Bot. Gard., 80, 902, 1993.
Quadros, F.L.F. and Pillar, V.P. Transições floresta–campo no Rio Grande do Sul, Ciênc. Amb., 24, 109, 2002.
Rambo, B., A mata pluvial do Alto Uruguai, Roessléria 3, 101, 1980.
Ratter, J.A. et al., Observations on the vegetation of northeastern Mato Grosso, II. Forests and soils of the
Rio Suiá-Missu area. Proc. R. Soc. London B, 203, 191, 1978.
Ratter, J.A. et al., Analysis of the floristic composition of the Brazilian cerrado vegetation. Edinburgh J. Bot.,
53, 153, 1996.
2987_C007.fm Page 179 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
179
Ribeiro, J.E.L.S. et al., Flora da Reserva Ducke: Guia de Identificação das Plantas Vasculares de uma Floresta
de Terra-Firme na Amazônia Central. INPA-DFID, Manaus, 1999.
Rocha, P.L.B., Queiroz, L.P., and Pirani, J.R., Plant species and habitat structure in a sand dune field in the
Brazilian Caatinga: a homogeneous habitat harbouring an endemic biota, Rev. Brasil. Bot., 27, 739,
2004.
Rodal, M.J.N., Montane forests in Northeast Brazil: a phytogeographical approach, Bot. Jarh. Syst., 124, 1,
2002.
Rodal, M.J.N. and Nascimento, L.M., Levantamento florístico de uma floresta serrana da Reserva Biológica
de Serra Negra, microrregião de Itaparica, Pernambuco, Brasil, Acta Bot. Brasil., 16, 481, 2002.
Rodal, M.J.N. and Sampaio, E.V.S.B., A vegetação do Bioma Caatinga, in Vegetação eFflora da Caatinga,
Sampaio, E.V.S.B. et al., Eds., Associação Plantas do Nordeste, Recife, 2002, 11.
Roderjan, C.V. et al., As unidades fitogeográficas do Estado do Paraná, Ciênc. Amb., 24, 75, 2002.
Salgado-Labouriau, M.L. et al., Late Quaternary vegetational and climatic changes in cerrado and palm swamp
from central Brazil, Palaeogeogr. Palaeoclimat. Palaeocecol., 128, 215, 1997.
Salis, S.M., Shepherd, G.J., and Joly, C.A., Floristic comparison of mesophytic semi-deciduous forests of the
interior of the state of São Paulo, southeast Brazil, Vegetatio, 119, 155, 1995.
Scudeller V.V., Martins F.R., and Shepherd G.J., Distribution and abundance of arboreal species in the Atlantic
ombrophilous dense forest in Southeastern Brazil, Plant Ecol., 152, 185, 2001.
Silva, J.M.C. and Tabarelli, M. Tree species impoverishment and the future flora of the Atlantic Forest of
northeast Brazil, Nature, 404, 72, 2000.
Silva, J.M.C. and Tabarelli, M., The future of Atlantic forest in Northeastern Brazil, Conserv. Biol. 15, 819,
2001.
Spichiger, R. et al., Origin, affinities and diversity hotspots of the Paraguayan dendrofloras, Candollea, 50,
515, 1995.
Spichiger, R., Calange, C., and Bise, B., Geographical zonation in the Neotropics of tree species characteristic
of the Paraguay-Paraná Basin, J. Biogeogr. 31, 1489, 2004.
ter Steege, H. et al., An analysis of the floristic composition and diversity of Amazonian forests including
those of the Guiana Shield. J. Trop. Ecol., 16, 801, 2000.
Thornthwaite, C.W., An approach toward a rational classification of climate, Geogr. Rev., 38, 55, 1948.
Torres, R.B., Martins, F.R., and Gouvea, L.S.K., Climate, soil and tree flora relationships in forests in the
state of São Paulo, southeastern Brazil, Revta. Brasil. Bot., 20, 41, 1997.
Veloso, H.P., Rangel Filho, A.L.R., and Lima, J.C.A., Classificação da Vegetação Brasileira, Adaptada a um
Sistema Universal, Instituto Brasileiro de Geografia e Estatística, Rio de Janeiro, 1991.
Walter, H. Vegetation of the Earth and Ecological Systems of the Geo-Biosphere, 3rd ed., Springer-Verlag,
Berlin, 1985.
Zappi, D.C. et al., Lista das plantas vasculares de Catolés, Chapada Diamantina, Bahia, Brasil. Bol. Bot. Univ.
São Paulo, 21, 345, 2003.
APPENDIX. MOST FREQUENT SPECIES (>70% OF CHECKLISTS) IN
THE TREE FLORA OF SELECTED SDTF AND SDSF FORMATIONS
OF EASTERN SOUTH AMERICA.
Low altitude tropical seasonal forests — North-east region: Abarema cochliacarpos, Acacia
polyphylla, Albizia pedicellaris, A. polycephala, Allophylus edulis, Alseis pickelii, Anacardium
occidentale, Andira fraxinifolia, A. nitida, Apeiba tibourbou, Apuleia leiocarpa, Aspidosperma
pyrifolium, Astronium fraxinifolium, Bowdichia virgilioides, Brosimum gaudichaudii, B. guianense, Buchenavia capitata, Byrsonima sericea, Caesalpinia ferrea, Campomanesia aromatica, C.
dichotoma, Capparis flexuosa, Casearia sylvestris, Cecropia pachystachya, C. palmata, Cereus
jamacaru, Chamaecrista apoucouita, C. ensiformis, Chrysophyllum rufum, Clusia nemorosa, Coccoloba alnifolia, C. cordifolia, Cordia trichotoma, Coutarea hexandra, Cupania revoluta, Curatella
americana, Enterolobium contortisiliquum, Erythrina velutina, Erythroxylum citrifolium, Eschweilera
ovata, Eugenia florida, E. punicifolia, E. uniflora, Guapira noxia, G. opposita, G. pernambucensis,
Guarea guidonia, Guazuma ulmifolia, Guettarda platypoda, Himatanthus phagedaenicus, Hirtella
ciliata, H. racemosa, Hymenaea courbaril, H. rubriflora, Inga capitata, I. ingoides, I. laurina,
2987_C007.fm Page 180 Thursday, December 1, 2005 7:03 PM
180
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
I. thibaudiana, Lecythis pisonis, Luehea ochrophylla, L. paniculata, Manihot epruinosa, Manilkara
salzmannii, Maytenus distichophylla, M. erythroxylon, Miconia albicans, Myrcia multiflora, M.
sylvatica, M. tomentosa, Myrsine guianensis, Ocotea notata, Ouratea hexasperma, Palicourea crocea, Pera glabrata, Pogonophora schomburgkiana, Pouteria grandiflora, Pradosia lactescens, Protium heptaphyllum, Psidium oligospermum, Pterocarpus rohrii, Rauvolfia ligustrina, Sacoglottis
mattogrossensis, Schefflera morototoni, Spondias mombin, Strychnos parvifolia, Stryphnodendron
pulcherrimum, Swartzia pickelii, Tabebuia impetiginosa, T. roseo-alba, T. serratifolia, Talisia esculenta, Tapirira guianensis, Thyrsodium spruceanum, Trema micrantha, Vismia guianensis, Vitex
triflora, Ximenia americana, Ziziphus joazeiro, Zollernia latifolia.
High altitude tropical seasonal forests — North-east region: Acacia polyphylla, A. riparia, A.
tenuifolia, Albizia polycephala, Allophylus edulis, Anadenanthera colubrina, Bowdichia virgilioides, Buchenavia capitata, Byrsonima sericea, Caesalpinia ferrea, Campomanesia aromatica, Capparis flexuosa, Casearia sylvestris, Ceiba glaziovii, Celtis iguanaea, Clusia nemorosa, Copaifera
langsdorffii, Cordia trichotoma, Coutarea hexandra, Croton rhamnifolius, Cupania revoluta, Cyathea
microdonta, Enterolobium contortisiliquum, Erythroxylum citrifolium, Eugenia punicifolia, Guapira laxiflora, G. opposita, Guazuma ulmifolia, Guettarda sericea, Hymenaea courbaril, Machaerium
hirtum, Manilkara rufula, Maprounea guianensis, Maytenus obtusifolia, Miconia albicans, Myrcia
fallax, M. multiflora, M. sylvatica, M. tomentosa, Myroxylon peruiferum, Myrsine guianensis,
Ocotea duckei, Piptadenia stipulacea, Platymiscium floribundum, Prockia crucis, Psidium
guineense, Randia nitida, Roupala cearensis, Ruprechtia laxiflora, Sapium glandulosum, Schoepfia
brasiliensis, Senna macranthera, S. spectabilis, S. splendida, Tabebuia impetiginosa, T. serratifolia,
Talisia esculenta, Vitex rufescens, Zanthoxylum rhoifolium.
Low altitude tropical seasonal forests — East region: Acacia polyphylla, Aegiphila sellowiana,
Albizia polycephala, Alchornea glandulosa, Allophylus edulis, Amaioua guianensis, Anadenanthera
colubrina, Andira fraxinifolia, Aparisthmium cordatum, Apuleia leiocarpa, Aspidosperma parvifolium, Astrocaryum aculeatissimum, Astronium graveolens, Bathysa nicholsonii, Bauhinia fusconervis, Brosimum guianense, B. lactescens, Byrsonima sericea, Cabralea canjerana, Carpotroche
brasiliensis, Casearia sylvestris, C. ulmifolia, Cassia ferruginea, Cecropia glaziovii, C. hololeuca,
Cedrela fissilis, Copaifera langsdorffii, Croton urucurana, Cyathea delgadii, Dalbergia nigra, Endlicheria paniculata, Erythrina verna, Erythroxylum pelleterianum, E. pulchrum, Eugenia florida,
Euterpe edulis, Ficus gomelleira, Gallesia integrifolia, Guapira opposita, Guarea macrophylla,
Guatteria australis, G. villosissima, Guettarda uruguensis, Himatanthus lancifolius, Hortia arborea,
Hymenolobium janeirense, Inga capitata, I. vera, Joannesia princeps, Lacistema pubescens, Lecythis
lurida, L. pisonis, Luehea divaricata, L. grandiflora, Mabea fistulifera, Machaerium brasiliense, M.
hirtum, M. stipitatum, Maclura tinctoria, Maprounea guianensis, Melanoxylon brauna, Miconia
cinnamomifolia, Myrcia fallax, M. rufula, Myrciaria floribunda, Nectandra oppositifolia, Ocotea
dispersa, Pera glabrata, Piptadenia gonoacantha, Plathymenia reticulata, Platymiscium floribundum,
Platypodium elegans, Pogonophora schomburgkiana, Pourouma guianensis, Protium warmingianum, Pseudobombax grandiflorum, Pseudopiptadenia contorta, Pterocarpus rohrii, Pterygota
brasiliensis, Rollinia laurifolia, Senna macranthera, S. multijuga, Siparuna guianensis, Sorocea
guilleminiana, Sparattosperma leucanthum, Stryphnodendron pulcherrimum, Swartzia acutifolia,
S. myrtifolia, Syagrus romanzoffiana, Tabebuia serratifolia, Tabernaemontana hystrix, Tapirira
guianensis, Trichilia lepidota, T. pallida, Urbanodendron verrucosum, Virola bicuhyba, Vismia
guianensis, Xylopia brasiliensis, Xylopia sericea, Zanthoxylum rhoifolium.
High altitude tropical seasonal forests — East region: Aegiphila sellowiana, Alchornea triplinervia,
Amaioua guianensis, Anadenanthera colubrina, Andira fraxinifolia, Apuleia leiocarpa, Aspidosperma
discolor, A. olivaceum, Astronium graveolens, Bauhinia longifolia, Blepharocalyx salicifolius,
Bowdichia virgilioides, Byrsonima sericea, Cabralea canjerana, Campomanesia xanthocarpa, Casearia
arborea, C. decandra, C. obliqua, C. sylvestris, Cassia ferruginea, Cecropia glaziovii, C. hololeuca,
2987_C007.fm Page 181 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
181
C. pachystachya, Cedrela fissilis, Celtis iguanaea, Chrysophyllum gonocarpum, Clethra scabra,
Copaifera langsdorffii, Cordia sellowiana, Croton floribundus, C. urucurana, Cupania paniculata,
C. vernalis, Cyathea corcovadensis, C. delgadii, C. phalerata, Dalbergia frutescens, Dictyoloma
vandellianum, Eugenia florida, Gochnatia polymorpha, Guapira opposita, Guarea macrophylla,
Guatteria australis, G. sellowiana, G. villosissima, Guazuma ulmifolia, Hyptidendron asperrimum,
Inga laurina, I. marginata, I. sessilis, Kielmeyera lathrophyton, Lamanonia ternata, Leandra melastomoides, Luehea divaricata, Machaerium brasiliense, M. hirtum, M. nictitans, M. villosum,
Maprounea guianensis, Matayba elaeagnoides, Maytenus salicifolia, Miconia cinnamomifolia, M.
ligustroides, M. pepericarpa, Myrcia detergens, M. fallax, M. guianensis, M. rostrata, M. tomentosa,
Myrsine coriacea, M. umbellata, Nectandra lanceolata, N. oppositifolia, Ocotea corymbosa, O.
odorifera, O. spixiana, Pera glabrata, Platypodium elegans, Protium heptaphyllum, Prunus myrtifolia, Psychotria vellosiana, Rollinia laurifolia, Rollinia sylvatica, Roupala brasiliensis, Sapium
glandulosum, Sclerolobium rugosum, Senna macranthera, S. multijuga, Siparuna guianensis, Siphoneugena densiflora, Sorocea guilleminiana, Tabebuia serratifolia, Tapirira guianensis, T. obtusa,
Terminalia glabrescens, Tibouchina candolleana, Trichilia pallida, Vitex polygama, Vochysia tucanorum, Zanthoxylum rhoifolium.
Low altitude tropical seasonal forests — South-east region: Acacia polyphylla, Actinostemon
klotzschii, Aegiphila sellowiana, Albizia niopoides, Alchornea glandulosa, A. triplinervia, Allophylus edulis, Aloysia virgata, Annona cacans, Apuleia leiocarpa, Aralia warmingiana, Aspidosperma
polyneuron, Astronium graveolens, Balfourodendron riedelianum, Bastardiopsis densiflora,
Cabralea canjerana, Campomanesia guazumifolia, C. xanthocarpa, Cariniana estrellensis, Casearia
gossypiosperma, C. sylvestris, Cecropia pachystachya, Cedrela fissilis, Ceiba speciosa, Celtis iguanaea, Chrysophyllum gonocarpum, C. marginatum, Colubrina glandulosa, Copaifera langsdorffii,
Cordia ecalyculata, C. trichotoma, Croton floribundus, Cupania vernalis, Dalbergia frutescens,
Dendropanax cuneatus, Diatenopteryx sorbifolia, Endlicheria paniculata, Enterolobium contortisiliquum, Esenbeckia febrifuga, Eugenia florida, E. involucrata, Euterpe edulis, Gallesia integrifolia,
Guapira opposita, Guarea guidonia, G. kunthiana, G. macrophylla, Guatteria australis, Gymnanthes
concolor, Heliocarpus americanus, Holocalyx balansae, Inga marginata, I. striata, I. vera, Ixora
venulosa, Jacaranda micrantha, Jacaratia spinosa, Lonchocarpus cultratus, L. muehlbergianus, Luehea divaricata, Machaerium hirtum, M. nictitans, M. paraguariense, M. stipitatum, Maclura tinctoria,
Matayba elaeagnoides, Metrodorea nigra, Myrcia multiflora, Myrciaria floribunda, Myrocarpus
frondosus, Myrsine umbellata, Nectandra megapotamica, Ocotea diospyrifolia, O. puberula, O.
pulchella, Parapiptadenia rigida, Patagonula americana, Peltophorum dubium, Piper amalago,
Prunus myrtifolia, Rollinia emarginata, R. sylvatica, Roupala brasiliensis, Schefflera morototoni,
Sebastiania commersoniana, Seguieria langsdorffii, Sorocea bonplandii, Syagrus romanzoffiana,
Tabernaemontana catharinensis, Terminalia triflora, Trema micrantha, Trichilia catigua, T. clausseni,
T. elegans, T. pallida, Vitex megapotamica, Zanthoxylum caribaeum, Z. fagara, Z. rhoifolium,
Z. riedelianum.
High altitude tropical seasonal forests — South-east region: Aegiphila sellowiana, Albizia
polycephala, Alchornea triplinervia, Amaioua guianensis, Andira fraxinifolia, Annona cacans, Aspidosperma olivaceum, Byrsonima laxiflora, Cabralea canjerana, Calyptranthes clusiifolia, Campomanesia guazumifolia, Cariniana estrellensis, Casearia decandra, C. lasiophylla, C. obliqua, C.
sylvestris, Cecropia glaziovii, C. pachystachya, Cedrela fissilis, Chrysophyllum marginatum, Cinnamomum glaziovii, Clethra scabra, Copaifera langsdorffii, Cordia sellowiana, Croton floribundus,
C. verrucosus, Cryptocarya aschersoniana, Cupania vernalis, Cyathea delgadii, C. phalerata, Dalbergia villosa, Daphnopsis brasiliensis, D. fasciculata, Dendropanax cuneatus, Endlicheria paniculata, Eugenia florida, Gomidesia affinis, Guapira opposita, Guarea macrophylla, Guatteria australis,
Gymnanthes concolor, Heisteria silvianii, Hyeronima ferruginea, Inga striata, Ixora warmingii,
Jacaranda macrantha, Lamanonia ternata, Leucochloron incuriale, Lithraea molleoides, Luehea
divaricata, L. grandiflora, Machaerium hirtum, M. nictitans, M. stipitatum, M. villosum,
2987_C007.fm Page 182 Thursday, December 1, 2005 7:03 PM
182
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Maclura tinctoria, Matayba elaeagnoides, Miconia cinnamomifolia, Mollinedia widgrenii, Myrcia
fallax, M. rostrata, Myrciaria floribunda, Myrsine coriacea, M. umbellata, Nectandra grandiflora,
N. oppositifolia, Ocotea corymbosa, O. diospyrifolia, O. odorifera, O. pulchella, Pera glabrata,
Persea pyrifolia, Piptocarpha macropoda, Platycyamus regnellii, Platypodium elegans, Protium
widgrenii, Prunus myrtifolia, Psychotria vellosiana, Rollinia dolabripetala, R. laurifolia, R. sylvatica, Roupala brasiliensis, Sapium glandulosum, Schinus terebinthifolius, Sclerolobium rugosum,
Solanum pseudoquina, Sorocea bonplandii, Syagrus romanzoffiana, Tabebuia serratifolia, Tapirira
guianensis, T. obtusa, Ternstroemia brasiliensis, Trichilia emarginata, T. pallida, Vernonanthura
diffusa, Vismia brasiliensis, Vitex polygama, Vochysia tucanorum, Xylopia brasiliensis, Zanthoxylum rhoifolium.
Low altitude tropical seasonal forests — Central-west region: Acacia polyphylla, Acrocomia
aculeata, Albizia niopoides, Alibertia concolor, Anadenanthera colubrina, A. peregrina, Apeiba
tibourbou, Apuleia leiocarpa, Aspidosperma cylindrocarpon, A. olivaceum, A. pyrifolium, A. subincanum, Astronium fraxinifolium, Attalea phalerata, Bauhinia longifolia, Bowdichia virgilioides,
Cabralea canjerana, Callisthene fasciculata, C. major, Calophyllum brasiliense, Cariniana estrellensis, Casearia gossypiosperma, C. rupestris, C. sylvestris, Cecropia pachystachya, Cedrela fissilis,
Ceiba speciosa, Celtis iguanaea, Chrysophyllum gonocarpum, Combretum leprosum, Copaifera
langsdorffii, Cordia glabrata, C. trichotoma, Coutarea hexandra, Cupania vernalis, Dilodendron
bipinnatum, Diospyros hispida, D. sericea, Enterolobium contortisiliquum, Eugenia florida, Genipa
americana, Guapira opposita, Guarea guidonia, Guazuma ulmifolia, Guettarda uruguensis, Hymenaea courbaril, Inga laurina, I. marginata, I. vera, Jacaranda cuspidifolia, Licania apetala, Luehea
divaricata, L. paniculata, Machaerium hirtum, M. stipitatum, M. villosum, Maclura tinctoria, Magonia pubescens, Matayba guianensis, Micropholis venulosa, Myracrodruon urundeuva, Myrcia
tomentosa, Myrciaria floribunda, Plathymenia reticulata, Platypodium elegans, Pouteria gardneri,
Protium heptaphyllum, P. spruceanum, Pseudobombax tomentosum, Psidium guineense, Pterogyne
nitens, Qualea multiflora, Randia nitida, Rhamnidium elaeocarpum, Rollinia emarginata, Salacia
elliptica, Sapium glandulosum, Sclerolobium paniculatum, Simira sampaioana, Siparuna guianensis, Sorocea guilleminiana, Spondias mombin, Sterculia striata, Sweetia fruticosa, Tabebuia impetiginosa, T. roseo-alba, T. serratifolia, Talisia esculenta, Tapirira guianensis, Terminalia argentea,
T. glabrescens, Tocoyena formosa, Trichilia catigua, T. clausseni, T. elegans, T. pallida, Triplaris
gardneriana, Unonopsis lindmanii, Vitex cymosa, Zanthoxylum rhoifolium.
High altitude tropical seasonal forests — Central-west region: Aegiphila sellowiana, Alibertia
edulis, Amaioua guianensis, Anadenanthera colubrina, Apuleia leiocarpa, Aspidosperma cylindrocarpon, A. australe, A. subincanum, Astronium fraxinifolium, Bauhinia longifolia, Cabralea canjerana, Callisthene major, Calophyllum brasiliense, Cardiopetalum calophyllum, Cariniana estrellensis, Casearia sylvestris, Cecropia pachystachya, Cedrela fissilis, Cheiloclinium cognatum,
Chrysophyllum marginatum, Copaifera langsdorffii, Cordia sellowiana, C. trichotoma, Cryptocarya
aschersoniana, Cupania vernalis, Dendropanax cuneatus, Diospyros hispida, Emmotum nitens,
Endlicheria paniculata, Eugenia florida, Euplassa inaequalis, Faramea cyanea, Ferdinandusa speciosa, Gomidesia fenzliana, Guarea guidonia, G. macrophylla, Guatteria sellowiana, Guazuma
ulmifolia, Guettarda uruguensis, Hedyosmum brasiliense, Hirtella glandulosa, H. gracilipes,
Hyeronima alchorneoides, Hymenaea courbaril, Inga alba, I. laurina, I. vera, Ixora warmingii,
Lamanonia ternata, Licania apetala, Luehea divaricata, Machaerium acutifolium, Maprounea guianensis, Matayba guianensis, Mauritia flexuosa, Miconia chamissois, Micropholis venulosa, Mouriri
glazioviana, Myrcia rostrata, M. tomentosa, Myrsine guianensis, M. umbellata, Nectandra cissiflora,
Ocotea corymbosa, O. spixiana, Ormosia fastigiata, Ouratea castaneifolia, Pera glabrata, Piptadenia
gonoacantha, Piptocarpha macropoda, Platypodium elegans, Pouteria gardneri, P. ramiflora, Protium
heptaphyllum, P. spruceanum, Prunus myrtifolia, Pseudolmedia laevigata, Qualea dichotoma, Q.
multiflora, Richeria grandis, Roupala brasiliensis, Schefflera morototoni, Sclerolobium paniculatum, Siparuna guianensis, Siphoneugena densiflora, Styrax camporum, Symplocos nitens, Tabebuia
2987_C007.fm Page 183 Thursday, December 1, 2005 7:03 PM
Floristic Relationships of Seasonally Dry Forests of Eastern
183
serratifolia, Talauma ovata, Tapirira guianensis, Terminalia argentea, T. glabrescens, Trichilia
catigua, Virola sebifera, Vitex polygama, Vochysia tucanorum, Xylopia aromatica, Xylopia emarginata, X. sericea, Zanthoxylum rhoifolium.
Subtropical seasonal forests — South region: Aiouea saligna, Alchornea triplinervia, Allophylus
edulis, A. guaraniticus, Apuleia leiocarpa, Banara parviflora, B. tomentosa, Blepharocalyx salicifolius, Cabralea canjerana, Calyptranthes concinna, Campomanesia xanthocarpa, Casearia decandra, C. sylvestris, Cedrela fissilis, Celtis iguanaea, Chrysophyllum gonocarpum, C. marginatum,
Citronella paniculata, Cordia ecalyculata, C. trichotoma, Cupania vernalis, Dalbergia frutescens,
Daphnopsis racemosa, Dasyphyllum spinescens, Diospyros inconstans, Enterolobium contortisiliquum, Erythrina crista-galli, Erythroxylum argentinum, Eugenia hyemalis, E. involucrata, E. opaca,
E. ramboi, E. rostrifolia, E. uniflora, Ficus insipida, F. luschnathiana, F. organensis, Gomidesia
palustris, Guapira opposita, Guettarda uruguensis, Gymnanthes concolor, Helietta apiculata, Ilex
brevicuspis, Inga marginata, I. vera, Jacaranda micrantha, Lithraea brasiliensis, Lonchocarpus
nitidus, Luehea divaricata, Machaerium stipitatum, Matayba elaeagnoides, Maytenus ilicifolia,
Myrcianthes pungens, Myrciaria tenella, Myrocarpus frondosus, Myrsine coriacea, M. loefgrenii,
M. lorentziana, M. umbellata, Nectandra lanceolata, N. megapotamica, Ocotea puberula, O. pulchella, Parapiptadenia rigida, Patagonula americana, Phytolacca dioica, Pilocarpus pennatifolius,
Pisonia zapallo, Pouteria gardneriana, P. salicifolia, Prunus myrtifolia, P. subcoriacea, Psidium
cattleianum, Quillaja brasiliensis, Randia nitida, Rollinia emarginata, R. sylvatica, Ruprechtia
laxiflora, Sapium glandulosum, Schefflera morototoni, Schinus terebinthifolius, Sebastiania brasiliensis, S. commersoniana, Seguieria americana, Solanum granuloso-leprosum, S. pseudoquina, S.
sanctaecatharinae, Sorocea bonplandii, Strychnos brasiliensis, Styrax leprosus, Syagrus romanzoffiana, Terminalia australis, Trema micrantha, Trichilia clausseni, T. elegans, Urera baccifera, Vitex
megapotamica, Xylosma pseudosalzmanii, Zanthoxylum fagara, Z. rhoifolium.
Subtropical seasonal forests — Southwest region: Acacia albicorticata, A. caven, A. praecox,
Acanthosyris falcata, Achatocarpus praecox, Allophylus edulis, Amburana cearensis, Anadenanthera colubrina, Aralia angelicifolia, Aspidosperma olivaceum, A quebracho-blanco, Caesalpinia
paraguariensis, Calycophyllum multiflorum, Capparis retusa, Carica quercifolia, Casearia sylvestris,
Celtis pubescens, Chloroleucon tenuiflorum, Chrysophyllum gonocarpum, C. marginatum, Cochlospermum tetraporum, Cordia trichotoma, Crataeva tapia, Cupania vernalis, Diplokeleba floribunda,
Enterolobium contortisiliquum, Erythrina falcata, Eugenia uniflora, Geoffroea decorticans, G. striata, Gleditsia amorphoides, Holocalyx balansae, Maclura tinctoria, Myracrodruon balansae, M.
urundeuva, Myrcianthes cisplatensis, M. pungens, Myrsine laetevirens, Parkinsonia aculeata, Patagonula americana, Peltophorum dubium, Phyllostylon rhamnoides, Phytolacca dioica, Pilocarpus
pennatifolius, Pisonia aculeata, P. zapallo, Pouteria gardneriana, Prosopis nigra, Pterogyne nitens,
Rollinia emarginata, Ruprechtia laxiflora, Sapindus saponaria, Sapium haematospermum, Schinopsis brasiliensis, Schinus polygamus, Scutia buxifolia, Sebastiania brasiliensis, Sideroxylon obtusifolium, Solanum granuloso-leprosum, Syagrus romanzoffiana, Tabebuia heptaphylla, T. impetiginosa, Tabernaemontana catharinensis, Terminalia triflora, Tipuana tipu, Ximenia americana,
Zanthoxylum fagara, Z. petiolare, Z. rhoifolium, Ziziphus mistol.
Chaco forests: Acacia aroma, A. caven, A. curvifructa, A. furcatispina, A. praecox, A. tucumanensis,
Acanthosyris falcata, Achatocarpus praecox, Allophylus edulis, Anadenanthera colubrina, Aporosella chacoensis, Aralia angelicifolia, Aspidosperma quebracho-blanco, A triternatum, Athyana
weinmanniifolia, Bougainvillea campanulata, B. praecox, Bulnesia bonariensis, Caesalpinia paraguariensis, Capparis atamisquea, C. retusa, C. salicifolia, C. insignis, Cereus stenogonus, Chrysophyllum marginatum, Copernicia alba, Cupania vernalis, Enterolobium contortisiliquum, Eugenia
uniflora, Geoffroea decorticans, Jacaratia corumbensis, Maytenus scutioides, M. spinosa, M. vitisidaea, Mimosa castanoclada, M. chacoensis, M. detinens, M. glutinosa, Mimozyganthus carinatus,
Myrcianthes cisplatensis, M. pungens, Myrsine laetevirens, Parkinsonia praecox, Patagonula americana,
2987_C007.fm Page 184 Thursday, December 1, 2005 7:03 PM
184
Neotropical Savannas and Dry Forests: Diversity, Biogeography, and Conservation
Pereskia sacharosa, Phyllostylon rhamnoides, Pisonia zapallo, Prosopis affinis, P. alpataco, P. elata,
P. fiebrigii, P. kuntzei, P. nigra, P. nuda, P. rojasiana, P. ruscifolia, P. sericantha, P. torquata,
Quiabentia chacoensis, Ruprechtia apetala, R. laxiflora, R. triflora, Sapium haematospermum,
Schinopsis balansae, S. cornuta, S. heterophylla, S. quebracho-colorado, Schinus polygamus, Scutia
buxifolia, Sesbania virgata, Sideroxylon obtusifolium, Stetsonia coryne, Tabebuia impetiginosa,
T nodosa, Tessaria dodoneifolia, T integrifolia, Thevetia bicornuta, Trema micrantha, Trithrinax
schizophylla, Ximenia americana, Zanthoxylum coco, Z. petiolare, Ziziphus mistol.
Caatingas: Acacia langsdorffii, A. paniculata, A. polyphylla, Allophylus quercifolius, Amburana
cearensis, Anadenanthera colubrina, Annona spinescens, Aspidosperma pyrifolium, Auxemma
glazioviana, A. oncocalyx, Balfourodendron molle, Bauhinia acuruana, B. cheilantha, B. pentandra,
Bocoa mollis, Brasiliopuntia brasiliensis, Byrsonima gardneriana, Caesalpinia bracteosa, C. ferrea,
C. microphylla, C. pyramidalis, Capparis flexuosa, C. jacobinae, C. yco, Ceiba glaziovii, Cereus
albicaulis, C. jamacaru, Chloroleucon acacioides, C. foliolosum, C. mangense, Cnidoscolus
bahianus, C. obtusifolius, C. quercifolius, Cochlospermum vitifolium, Combretum leprosum, Commiphora leptophloeos, Cordia leucocephala, Coutarea hexandra, Croton rhamnifolius, C. sonderianus, Dalbergia catingicola, D. cearensis, Erythrina velutina, Erythroxylum revolutum, Eugenia
punicifolia, E. tapacumensis, Fraunhoffera multiflora, Geoffroea spinosa, Guapira laxa, Jatropha
mollissima, J. mutabilis, Manihot dichotoma, M. glaziovii, Maytenus rigida, Mimosa arenosa, M.
caesalpinifolia, M. gemmulata, M. malacocentra, M. tenuiflora, Myracrodruon urundeuva,
Parapiptadenia zehntneri, Pilosocereus gounellei, P. pachycladus, P. tuberculatus, Piptadenia obliqua, P. stipulacea, Pithecellobium diversifolium, Pseudobombax simplicifolium, Rollinia leptopetala, Sapium argutum, Schinopsis brasiliensis, Senna acuruensis, S. spectabilis, Senna splendida,
Sideroxylon obtusifolium, Spondias tuberosa, Tabebuia impetiginosa, Tacinga inamoena, T. palmadora, Thiloa glaucocarpa, Zanthoxylum stelligerum, Ziziphus joazeiro.
SDTF ‘Supertramp’ species (present in >100 checklists): Acacia polyphylla, Acrocomia aculeata,
Aegiphila sellowiana, Alibertia concolor, Allophylus edulis, Aloysia virgata, Anadenanthera colubrina, Andira fraxinifolia, Apuleia leiocarpa, Aspidosperma olivaceum, A. pyrifolium, Astronium
fraxinifolium, Bauhinia forficata, Bowdichia virgilioides, Brosimum gaudichaudii, Cabralea canjerana, Campomanesia xanthocarpa, Casearia decandra, C. sylvestris, Cecropia pachystachya,
Cedrela fissilis, Ceiba speciosa, Celtis iguanaea, C. pubescens, Chrysophyllum gonocarpum, C.
marginatum, Copaifera langsdorffii, Cordia trichotoma, Coutarea hexandra, Cupania vernalis, Dalbergia frutescens, Diospyros inconstans, Endlicheria paniculata, Enterolobium contortisiliquum,
Eugenia florida, E. punicifolia, E. uniflora, Garcinia gardneriana, Guapira opposita, Guarea guidonia, Guazuma ulmifolia, Guettarda uruguensis, Gymnanthes concolor, Hymenaea courbaril, Inga
marginata, I. vera, Lithraea molleoides, Lonchocarpus campestris, Luehea divaricata, L. grandiflora,
Machaerium acutifolium, M. hirtum, M. stipitatum, Maclura tinctoria, Maprounea guianensis,
Matayba elaeagnoides, M. guianensis, Maytenus ilicifolia, Miconia albicans, Myracrodruon urundeuva, Myrcia guianensis, M. multiflora, M. rostrata, M. tomentosa, Myroxylon peruiferum, Peltophorum dubium, Pera glabrata, Piper amalago, Pisonia zapallo, Platypodium elegans, Prockia
crucis, Protium heptaphyllum, Prunus myrtifolia, Pterogyne nitens, Randia nitida, Rollinia emarginata, R. sylvatica, Roupala brasiliensis, Ruprechtia laxiflora, Sapium glandulosum, Schefflera
morototoni, Sebastiania brasiliensis, Sideroxylon obtusifolium, Siparuna guianensis, Solanum
granuloso-leprosum, Sweetia fruticosa, Syagrus oleracea, S. romanzoffiana, Tabebuia impetiginosa,
T. serratifolia, Tapirira guianensis, Terminalia fagifolia, Trema micrantha, Trichilia catigua, T. clausseni,
T. elegans, Urera baccifera, Zanthoxylum fagara, Z. petiolare, Z. rhoifolium.